content
stringlengths 1
15.9M
|
---|
\section{Introduction}
Problem of allocating resources among selfish agents has been a
well-established research theme in economics and recently becomes
an emerging research topic in AI because AI methodologies can
provide computational techniques \cite{Rothkopf,Sandholm,Huang} to
the balancing of computation tractability and economic (or
societal) needs in these problems.
Dynamic mechanisms for resource allocation are trading mechanisms
for discovering market-clearing prices and equilibrium allocations
based on price adjustment processes \cite{Ausubel,Gul1,Huang}
Assume a seller wishes to sell a set of indivisible items to a
number of buyers. The seller announces the current prices of the
items and the buyers respond by reporting the set of items they
wish to buy at the given prices. The seller then calculates the
over-demanded set of items and increases the prices of
over-demanded items. This iterative process continues until all
the selling items can be sold at the prices at which each buyer is
assigned with items that maximize her personal net benefit.
Different from one-shot combinatorial auctions\cite{Cramton}, the main issue of
a dynamic mechanism is whether the procedure can lead to an
equilibrium state (Walrasian equilibrium) at which all the selling
items are effectively allocated to the buyers (equilibrium
allocation) and the price of items gives the buyers their best
values \cite{Gul2,Kelso,Lehmann,Sun}.
Most of the discussions on the issues of dynamic mechanisms are
based on market models in which there does not exist price
rigidities. In fact, ``good" allocations must look after both
sides economy efficiency and social equality, and price rigidities
may play a key role in some of these problems. For instance, in an
estate bubble period, housing cost is unbearable for most of the
members of society. The government may need to allocate some
housing resources (whose prices are not completely flexible but
restricted under some price rigidities) to middle-income earners.
On one hand, the lower bound prices can be made according to some
basic economic requirements (e.g., construction costs); on the
other hand, the upper bound prices \footnote{Note that since upper
bound prices are often set for the sake of equality between social
members (who have some but limited pay ability), they generally
accompany a limit to the number of resources one member can get.}
should be made according to some realistic social foundation
(e.g., average income level or pay ability). It is well-known that
a Walarasian equilibrium exists in the economy when there are no
price rigidities. In the case of price restrictions, a Walrasian
equilibrium may not exist since the equilibrium price vector may
not be admissible.
Talman and Yang studied the equilibrium allocation of heterogeneous
indivisible items under price rigidities, and proposed the
concept of constrained Walrasian equilibria \cite{Talman}. A
constrained Walrasian equilibrium consists of a price vector
$\textbf{p}$, a rationing system $R$, and a (constrained)
equilibrium allocation $\pi$ \cite{Lehmann} s.t. $\textbf{p}$
obeys the price rigidities, and $\pi$ assigns each buyer an
item (permitted by $R$) that maximizes her personal net
benefit at $\textbf{p}$. They also proposed two dynamic
auction procedures that produce constrained
Walrasian equilibria. However, the computational issues of these procedures have not been touched.
In this paper, we present a polynomial algorithm that
can be used to find over-demanded sets of items, and then
introduce a dynamic mechanism(called MAPR) with rationing to discover
constrained Walrasian equilibria under price rigidities in
polynomial time. In
MAPR, buyers compete with each other (with the help of the
seller) on prices of items for multiple rounds. In each
round, the seller announces the current price vector
(initially, the lower bound price vector) of the items that
remain, then the buyers respond by reporting the set of resources
they wish to buy, then the seller computes a minimal
over-demanded set $X_{min}$ of the items. If
$X_{min}=\emptyset$ then the final allocation is computed by the
RM subroutine and MAPR stops. Otherwise if all the prices of the
items in $X_{min}$ are less than their upper bounds then the
seller increases them; else an item $a\in X_{min}$ (whose
price is on its upper bound) is picked and the buyers who only
demand some items (including $a$) in $X_{min}$ draw lots for
the right to buy $a$. Since MAPR's execution process is
nondeterministic, we define the concepts of buyers' expected
profits and items' expected prices, and consider strategical
issues (in the sense of expected profit) in MAPR.
Here are main contributions of our work:
\begin{itemize}
\item We address the computational problems of dynamic auction proposed by [Talman and Yang, 2008], where these problems have not been touched.
\item \cite{Talman} has not finished the proof about the existence of constrained Walrasian equilibrium. We propose an algorithm to get the final allocation and several lemmas to prove the criteria required in constrained Walrasian equilibrium.
\item We defined the ``expected profits'' and ``expected prices'' and discuss strategical issues.
\end{itemize}
This paper is structured as follows. First, we review some basic
notions that are relevant to our work (see \cite{Talman} for
further details and examples). Second, we represent demand
situations with bipartite graphs. Third, we address the
computation of minimal over--demanded sets of items. Fourth,
we present MAPR, and prove formally that it yields a constrained
Walrasian equilibrium in polynomial time. Fifth, we consider
strategical issues in MAPR. Finally, we draw some conclusions.
\section{Preliminaries}
Consider a market situation where a seller wishes to sell a finite set $X$ of
indivisible items to a finite
number of buyers
$N=\{1,2,\ldots,n\}$. The item $o\in X$ is a dummy item
which can be assigned to more than one buyer. Items (eg.,
houses or apartments) in $X\setminus\{o\}$ may be heterogeneous.
A price vector $\textbf{p}\in\mathbb{Z}_+^X$ assigns a
non-negative integer to each $a\in X$ and $\textbf{p}_a$ is the
price of $a$ under $\textbf{p}$. It is required that
$\textbf{p}_a$ is not completely flexible and restricted to an
interval $[\underline{\textbf{p}}_a, \overline{\textbf{p}}_a]$
s.t.
$\underline{\textbf{p}}_a,\overline{\textbf{p}}_a\in\mathbb{Z}_+$,
$\underline{\textbf{p}}_a\leq\overline{\textbf{p}}_a$, and
$0=\underline{\textbf{p}}_o=\overline{\textbf{p}}_o$. We say
$\underline{\textbf{p}}$ and $\overline{\textbf{p}}$ as the lower
and upper bound price vectors. $P=\{\textbf{p}\in
\mathbb{Z}_+^X|(\forall a\in
X)\underline{\textbf{p}}_a\leq\textbf{p}_a\leq\overline{\textbf{p}}_a\}$
is called the set of \emph{admissible} price vectors. Each $i\in
N$ has an integer value function, i.e., $u_i : X \rightarrow
\mathbb{Z}_+$. $u_i(a)$ is i's valuation to item $a$. We
assume $u_i$ is i's private information, $u_i(o) = 0$, and i can
pay $\max_{a\in X}\overline{\textbf{p}}_a$ units of money. We say
$E=\langle N,X,\{u_i\}_{i\in N}\rangle$ is an \textit{economy}.
A rationing system is a function $R:N\times X\rightarrow\{0,1\}$
s.t. $R(i,o)=1$ for every $i\in N$. $R(i,a)=1$ means that buyer
$i$ is allowed to demand item $a$, while $R(i,a)=0$ means that
$i$ is not allowed to demand $a$. At $\textbf{p}\in P$ and
rationing system $R$, the indirect utility $V_i(\textbf{p},R)$ and
constrained demand $D_i(\textbf{p},R)$ of buyer $i$ is given by:
$V_i(\textbf{p},R)=\max\{u_i(a)-\textbf{p}_a|a\in X\textrm{ and
}R(i,a)=1\}$, and $D_i(\textbf{p},R)=\{a\in X|R(i,a)=1\textrm{ and
}u_i(a)-\textbf{p}_a=V_i(\textbf{p},R)\}$. An \textit{allocation}
of $X$ is a function $\pi:N\rightarrow X$ s.t. $\pi(i)\neq \pi(j)$
if $j\neq i$ and $\pi(i)\in X\setminus\{o\}$. $\pi$ is an
\textit{equilibrium allocation} if $\pi(i)\in D_i(\textbf{p},R)$
for all $i\in N$.
$\langle \textbf{p},R,\pi\rangle$ is a \textit{constrained
Walrasian equilibrium} if \textbf{(1)} $\textbf{p}\in P$, $R$ is a
rationing system, \textbf{(2)} $\pi$ is an equilibrium allocation,
\textbf{(3)} $\textbf{p}_a=\underline{\textbf{p}}_a$ if
$\pi(i)\neq a$ for all $i\in N$, \textbf{(4)}
$\textbf{p}_a=\overline{\textbf{p}}_a$ and $\pi(i)=a$ for some
$i\in N$ if $R(j,a)=0$ for some $j\in N$, and \textbf{(5)} $a\in
D_i(\textbf{p},R')$ if $R(i,a)=0$, where $R'(j,b)=R(j,b)$ for all
$\langle j,b\rangle\in N\times X$ except $R'(i,a)=1$.
Conditions \textbf{(1)} and \textbf{(2)} need no explanation.
Condition \textbf{(3)} says that if the price of a item is greater
than its lower bound then it must be assigned to some buyer.
\textbf{(4)} states that if an buyer is not allowed to demand some
items then the item must be assigned to another buyer at its upper
bound price. Condition \textbf{(5)} says that if an buyer is
allowed to demand a item which she was not allowed to demand, then
she will demand the item. To sum up, \textit{constrained Walrasian
equilibrium} is a equilibrium state under price rigidities. All
the five conditions make a balance between efficiency and
equality.
The following example is modified from the one given in
\cite{Talman}. It illustrates the notions introduced in this
section and will be used throughout the paper.
\begin{Example}
\begin{small}
Let $E=\langle N,X,\{u_i\}_{i\in N}\rangle$ be an economy such
that $N=\{1,2,3,4,5\}$, $X=\{o,a,b,c,d\}$, and buyers' values are
given in Table 1; price vector $\textbf{p}=(0,5,4,4,7)$; and $\pi$
be an allocation of X such that $\pi(1)=o$, $\pi(2)=c$,
$\pi(3)=b$, $\pi(4)=a$, and $\pi(5)=d$. Suppose the lower and
upper bound price vectors are $\underline{\textbf{p}}=(0, 5, 4, 1,
5)$, and $\overline{\textbf{p}}=(0, 6, 6, 4, 7)$, respectively. So
$\textbf{p}$ is an admissible price vector. Let $R$ be a rationing
system such that $R(i,x)=1$ for all $\langle i,x\rangle\in N\times
X$ except that $R(3,c)=R(1,c)=0$. For each buyer $i\in N$,
$V_i(\textbf{p},R)$ and $D_i(\textbf{p},R)$ are also shown in
Table 1. Obviously, $\langle\textbf{p},R,\pi\rangle$ is a
constrained Walrasian equilibrium.
\end{small}
\end{Example}
\begin{table}
\caption{Values, Indirect Utilities, and Constrained Demand} \setlength\tabcolsep{3pt}
\begin{center}
\begin{footnotesize}
\begin{tabular}{c|c|c|c|c|c|c|c}
\hline
buyer i & $u_{i}(o)$ & $u_{i}(a)$ & $u_{i}(b)$ & $u_{i}(c)$ & $u_{i}(d)$ & $V_i(\textbf{p},R)$ & $D_i(\textbf{p},R)$\\
\hline
1 & 0 & 4 & 3 & 5 & 7 & 0 & \{o,d\}\\
2 & 0 & 7 & 6 & 8 & 3 & 4 & \{c\}\\
3 & 0 & 5 & 5 & 8 & 7 & 1 & \{b\}\\
4 & 0 & 9 & 4 & 3 & 2 & 4 & \{a\}\\
5 & 0 & 6 & 2 & 4 & 10 & 3 & \{d\}\\
\hline
\end{tabular}
\end{footnotesize}
\end{center}
\end{table}
\section{Demand Situation and Maximum Consistent Allocation}
Given an economy $E=\langle N,X,\{u_i\}_{i\in N}\rangle$, we call
$\mathcal{D} = (D_i)_{i\in N}$ a \textit{demand situation} of $E$
if there is a price vector $\textbf{p}$ and a rationing system $R$
such that $D_i = D_i(\textbf{p},R)$ for all $i\in N$. An
allocation $\pi$ is \emph{consistent} with $\mathcal{D}$ if
$\pi(i)\in D_i\cup\{o\}$ for all $i\in N$. $\pi$ is maximum if
$|\{i\in N|o\not\in D_i\; and\;\pi(i)\neq o\}|$ $\geq|\{i\in
N|o\not\in D_i\; and\;\pi'(i)\neq o\}|$ for every allocation
$\pi'$ consistent with $\mathcal{D}$.
$\mathcal{D}$ can be represented as a bipartite graph
$\textsc{BG}(\mathcal{D}) =\langle N'\cup X',\mathcal{E}\rangle$
where $N'=\{i\in N|o\not\in D_i\}$, $X'=\bigcup_{i\in N'}D_i$, and
$\mathcal{E}=\{\{i,a\}|i\in N',a\in D_i\}$. A \textit{matching} in
$\textsc{BG}(\mathcal{D})$ is a subset $M$ of $\mathcal{E}$ s.t.
$e\cap e'=\emptyset$ for all $e,e'\in M$ with $e\neq e'$. $M$ is
maximum if $|M'|\leq|M|$ for each matching $M'$.
It is not hard to see that a matching $M$ in
$\textsc{BG}(\mathcal{D})$ determines an allocation consistent
with $\mathcal{D}$. $\pi^M$ denotes the allocation determined by
$M$, that is, $\pi^M(i)=a$ if $\exists\{i,a\}\in M$, and
$\pi^M(i)=o$ otherwise. Suppose $M$ is maximum, then $\pi^M$ is
maximum and it is easy to find that: there exists an equilibrium
allocation $\Leftrightarrow$ $|M|=|\{i\in N|o\not\in D_i\}|$
$\Leftrightarrow$ $\pi^M$ is an equilibrium allocation.
In fact, to find a maximum matching in a bipartite graph is a pure
combinatorial optimization problem, which can be addressed in
polynomial time. \cite{Schrijver} presents the matching augmenting
algorithm $\textsc{Ma}$, which takes a bipartite graph
$\mathcal{G}=\langle \mathcal{V},\mathcal{E}\rangle$ and a
matching $M$ in $\mathcal{G}$ as input, and outputs a matching
$\textsc{Ma}(\mathcal{G},M)=M'$ s.t. $|M'|\geq|M|$ and
$\bigcup_{e\in M'}e\supseteq \bigcup_{e\in M}e$ in time
$O(|\mathcal{E}|)$. So a maximum matching can be found in time
$O(|\mathcal{V}||\mathcal{E}|)$ (as we do at most $|\mathcal{V}|$
iterations), i.e., $O(|N||X|\min(|N|,|X|))$. In the following
discussion, $\hat{M}_{\mathcal{D}}$ denotes the maximum matching
of $\textsc{BG}(\mathcal{D})$ found by this way.
\begin{Example}
\begin{small}
See the economy given in Example 1. Let price vector
$\textbf{p}=(0,5,4,3,5)$ and $R$ be the rationing system such that
$R(i,a)=1$ for all $\langle i,a\rangle\in N\times X$. Then buyers'
constrained demands at $\textbf{p}$ and $R$ are:
$D_1(\textbf{p},R)=\{c,d\}$,
$D_2(\textbf{p},R)=D_3(\textbf{p},R)=\{c\}$,
$D_4(\textbf{p},R)=\{a\}$, $D_5(\textbf{p},R)=\{d\}$. Let
$\mathcal{D}=(D_i(\textbf{p},R))_{i\in N}$. Then
$\hat{M}_{\mathcal{D}}=\{\{1,c\},\{4,a\},\{5,d\}\}$.
\end{small}
\end{Example}
\section{Over-demanded Set of Items}
What can lead to non-existence of equilibrium allocations? This is a
key issue that we need to consider.
Given a demand situation $\mathcal{D}=(D_i)_{i\in N}$, a set of
real items $X'\subseteq X\setminus\{o\}$ is
\textit{over-demanded} in $\mathcal{D}$, if the number of buyers
who demand only items in $X'$ is strictly greater than the
number of items in $X'$, i.e., $|\{i\in N|D_i\subseteq
X'\}|>|X'|$; $X'$ is \textit{not under-demanded}, if the number of
buyers who demand some items in $X'$ is not less than the
number of items in $X'$, i.e., $|\{i\in N|D_i\cap
X'\neq\emptyset\}|\geq|X'|$. An over-demanded set $X'$ is
\textit{minimal} if no strict subset of $X'$ is over-demanded. We
can get Lemma 1 directly based on these definitions.
\begin{Lemma}
Let $X'\subseteq X\setminus\{o\}$ is over-demanded. Then for each
$a\in X'$, either there exists a minimal over-demanded set
$X''\subseteq X'$ s.t. $a\not\in X''$, or $a\in X''$ for every
minimal over-demanded set $X''\subseteq X'$.
\end{Lemma}
Theorem 1 answers the question proposed in the beginning of this section.
\begin{Theorem}
There exists an over-demanded set of items in
$\mathcal{D}=(D_i)_{i\in N}$ if and only if there does not exist
an equilibrium allocation.
\end{Theorem}
\textsc{Proof.} \begin{small} Sufficiency is obvious. Let us prove
necessity. Suppose there does not exist an equilibrium allocation.
Let $M=\hat{M}_{\mathcal{D}}$ and $N'=\{i\in N|o\not\in D_i\}$.
Then $|M|=|N\cap\bigcup_{e\in M}e|<|N'|$. Pick a buyer $i$ from
$N'\setminus N\cap\bigcup_{e\in M}e$. We construct a sequence
$\langle X_0,N_0\rangle,\langle X_1,N_1\rangle,\ldots$ as follow:
\begin{itemize}
\item $X_0=D_i$, $N_0=\{j\in N|(\exists a\in X_0)\{j,a\}\in M\}$;
\item $X_{k+1}=\bigcup_{j\in N_k}D_j$; and $N_{k+1}=\{j\in
N|(\exists a\in X_{k+1})\{j,a\}\in M\}$.
\end{itemize}
Pick any $k\geq 0$ and $a\in X_{k}$. Suppose there does not exist
$j\in N$ such that $\{j,a\}\in M$. Then there is an $M$-augmenting
path \cite{Schrijver} from $a$ to $i$, i.e., $M$ is not maximum,
contradicting the fact that $M$ is maximum. So for all $k\geq 0$
and $a\in X_{k}$, there exists $j\in N$ such that $\{j,a\}\in M$.
Consequently,
\begin{enumerate}
\item $X_k\subseteq X_{k+1}\subseteq X$, $N_k\subseteq
N_{k+1}\subseteq N$ for all $k\geq 0$;
\item if $X_{k+1}=X_{k}$ then $X_{k+l}=X_{k}$ and $N_{k+l}=N_{k}$
for all $k,l\geq 0$.
\end{enumerate}
So there must exist $K\geq 0$ s.t. $X_{0}\subset\ldots\subset
X_{K}=X_{K+1}=\ldots$. For each $b\in X_{K}$, $b$ is assigned to
only one buyer in $N_{K}$ at $\pi^M$. And for each $j\in N_{K}$,
$D_j\subseteq X_{K}$ and $j$ is assigned with only one item in
$X_{K}$ at $\pi^M$. So $|X_{K}|=|N_{K}|$. Consequently, $|\{i\in
N|D_i\subseteq
X_{K}\}|\geq|N_{K}\cup\{i\}|=|N_{K}|+1=|X_{K}|+1>|X_{K}|$. So
$X_{K}$ is an over-demanded set of items in
$\mathcal{D}$.$\quad\square$
\end{small}
\begin{figure}
\begin{center}
\begin{small}
\begin{tabular}{rl}
\hline\noalign{\smallskip}
1. & $\underline{\textbf{algorithm}}$ $\textsc{MODS}(\mathcal{D}=(D_i)_{i\in N},M=\hat{M}_{\mathcal{D}})$\\
2. & $\quad$pick $i$ from $\{i\in N|o\not\in D_i\}\setminus\bigcup_{e\in M}e$;\\
3. & $\quad X'':=D_i,X':=\emptyset$;\\
4. & $\quad\underline{\textbf{while}}(X''\neq\emptyset)$\\
5. & $\qquad N':=\{j\in N|(\exists a\in X'')\{j,a\}\in M\}$;\\
6. & $\qquad X':=X'\cup X''$, $X'':=\bigcup_{j\in N'}D_j\setminus X'$;\\
7. & $\quad X_{min}:=\emptyset,X'':=X'$;\\
8. & $\quad\underline{\textbf{for}}$ all $a\in X'$\\
9. & $\qquad X'':=X''\setminus\{a\}$;\\
10. & $\qquad N':=\{i\in N|D_i\subseteq X_{min}\cup X''\}$;\\
11. & $\qquad\mathcal{D'}:=(D_i)_{i\in N'}$, $k:=|\hat{M}_{\mathcal{D'}}|$;\\
12. & $\qquad\underline{\textbf{if}}$ $k=|N'|$\\
13. & $\qquad\quad X_{min}:=X_{min}\cup\{a\}$;\\
14. & $\quad\underline{\textbf{return}}$ $X_{min}$;\\
\noalign{\smallskip} \hline
\end{tabular}
\end{small}
\caption{\textsc{MODS} algorithm.}
\end{center}
\end{figure}
To find a minimal over-demanded set of items, we develop the
\textsc{MODS} algorithm shown in Figure 1. Given a demand
situation $\mathcal{D}$, and $\hat{M}_{\mathcal{D}}$ s.t.
$|\hat{M}_{\mathcal{D}}|<|\{i\in N|o\not\in D_i\}|$,
$\textsc{MODS}$ returns a minimal over-demanded set of items
$X_{min}$. The basic idea of $\textsc{MODS}$ is to generate an
over-demanded set $X'$ firstly (see lines 2-6 in Figure 1), and
then (according to Lemma 1) to find a minimal over-demanded set
$X_{min}\subseteq X'$ (see lines 7-14 in Figure 1).
The correctness of algorithm \textsc{MODS} is directly from Lemma
1 and the proof of Theorem 1. Let $BG(\mathcal{D})=\langle
\mathcal{V,E}\rangle$. Observe \textsc{MODS} and we can find the
following facts.
\begin{enumerate}
\item In order to generate an over-demanded set $X'$ (lines 4-6 in
Figure 1), \textsc{MODS} only visits edges in $\mathcal{E}$. For
each $e\in\mathcal{E}$, $e$ can be visited once at most.
\item $|X'|\leq|\hat{M}_{\mathcal{D}}|\leq\min(|N|,|X|)$, and
$BG(\mathcal{D'})\subseteq BG(\mathcal{D})$ (see line 11).
\end{enumerate}
According to $|\mathcal{E}|\leq|N||X|$, and that the complexity of
$\hat{M}_{\mathcal{D}}$ is in $O(|N||X|\min(|N|,|X|))$, the
overall complexity of
$\textsc{MODS}(\mathcal{D},\hat{M}_{\mathcal{D}})$ is in
$O(|N||X|(\min(|N|,|X|))^2)$.
\begin{Example}
\begin{small}
See $\mathcal{D}$ and $\hat{M}_{\mathcal{D}}$ described in Example
2. It is easy to find that $|\hat{M}_{\mathcal{D}}|<|\{i\in N|
o\not\in D_i\}|$. We apply $\textsc{MODS}$ algorithm to
$(\mathcal{D},\hat{M}_{\mathcal{D}})$. Firstly, an over-demanded
set $X'=\{c,d\}$ is found. And then a minimal over-demanded set
$X_{min}=\{c\}$ is found.
\end{small}
\end{Example}
\section{Mechanism for Resource Allocation under Price Rigidities}
In this section, we present a polynomial mechanism for resource
allocation under price rigidities (MAPR). Its basic idea is to
eliminate over-demanded sets of items by increasing the prices
of over-demanded items or rationing an over-demanded item
whose price has reached its upper bound.
\begin{center}
\textsc{MAPR}
\end{center}
\begin{itemize}
\item[\textbf{(1)}] The seller $\varphi$ announces the set $X$
of items to allocate, and sets
$\textbf{p}^0:=\underline{\textbf{p}},M^0:=\emptyset,N':=N$. Each
buyer $i\in N$ sets $R_i[a]:=1$ for all $a\in X$. Let $t:=0$.
\item[\textbf{(2)}] $\varphi$ sends $\textbf{p}^t$ and ``Report
your demand." to each $i\in N'$.
\item[\textbf{(3)}] Each $i\in N'$ computes and sends
$D_i$\footnote{$D_i=\{a\in X|R_i[a]=1\textrm{ and
}u_i(a)-\textbf{p}^t_a=\max\{u_i(b)-\textbf{p}^t_b|R_i[b]=1\}\}$}
to $\varphi$.
\item[\textbf{(4)}] $\varphi$ computes $N''=\{i\in
N'|D_i\cap\bigcup_{e\in M^t}e\neq\emptyset\}$. If $N''=\emptyset$
then go to step (6). $\varphi$ sends ``Sorry, items in
$D'_i=D_i\cap\bigcup_{e\in M^t}e$ have been sold. Please report
your new demand.'' to each $i\in N''$, and sets $N':=N''$.
\item[\textbf{(5)}] Each $i\in N'$ sets $R_i[a]:=0$ for all $a\in
D'_i$. Go to (3).
\item[\textbf{(6)}] Let $N^*=N\setminus\bigcup_{e\in M^t}e$ and
$\mathcal{D}^*=(D_i)_{i\in N^*}$. $\varphi$ computes
$\hat{M}_{\mathcal{D}^*}$. If $|\hat{M}_{\mathcal{D}^*}|=|\{i\in
N^*|o\not\in D_i\}|$ then go to step (9). $\varphi$ computes
$X_{min}=\textsc{MODS}(\mathcal{D}^*,\hat{M}_{\mathcal{D}^*})$.
\item[\textbf{(7)}] $\varphi$ computes $\overline{X}=\{a\in
X_{min}|\textbf{p}^t_a=\overline{\textbf{p}}_a\}$. If
$\overline{X}=\emptyset$ then: $\varphi$ sets $N':=N^*$,
$M^{t+1}:=M^t$, $\textbf{p}^{t+1}_a:=\textbf{p}^{t}_a+1$ for all
$a\in X_{min}$, and $\textbf{p}^{t+1}_a:=\textbf{p}^{t}_a$ for all
$a\in X\setminus X_{min}$. Let t:=t+1. Go to (2).
\item[\textbf{(8)}] $\varphi$ picks an item $a$ from
$\overline{X}$ and asks the buyers in $\{i\in N^*|a\in
D_i\subseteq X_{min}\}$ to draw lots for the right to buy $a$. Let
$i$ be the winning buyer. $\varphi$ sets
$M^{t+1}:=M^t\cup\{\{i,a\}\}$, $N':=N^*\setminus\{i\}$ and
$\textbf{p}^{t+1}:=\textbf{p}^{t}$. Let t:=t+1. Go to (2).
\item[\textbf{(9)}] $\varphi$ computes
$M^*:=M^t\cup\textsc{RM}((D_i)_{i\in
N},M^t,\textbf{p}^t,\underline{\textbf{p}})$ and then announces
$\textbf{p}^{t}$ and $\pi^{M^*}$ are the final price vector and
allocation. MAPR stops.
\end{itemize}
\begin{figure}
\begin{center}
\begin{small}
\begin{tabular}{rl}
\hline\noalign{\smallskip}
1. & $\underline{\textbf{algorithm}}$ $\textsc{RM}((D_i)_{i\in N},M,\textbf{p},\underline{\textbf{p}})$\\
2. & $\quad X':=\{a\in X\setminus\bigcup_{e\in M}e|\textbf{p}_a>\underline{\textbf{p}}_a\}$;\\
3. & $\quad N':=\{i\in N\setminus\bigcup_{e\in M}e|D_i\cap X'\neq\emptyset\}$;\\
4. & $\quad \mathcal{D'}:=(D_i\cap X')_{i\in N'},\; M':=\hat{M}_{\mathcal{D'}}$;\\
5. & $\quad N^*:=N\setminus\bigcup_{e\in M}e$, $\langle\mathcal{V,E}\rangle:=BG((D_i)_{i\in N^*})$;\\
6. & $\quad M'':=M'\cap\mathcal{E}$;\\
7. & $\quad \underline{\textbf{while}}(\textsc{Ma}(\langle\mathcal{V,E}\rangle,M'')\neq M'')$\\
8. & $\qquad M'':=\textsc{Ma}(\langle\mathcal{V,E}\rangle,M'')$;\\
9. & $\quad \underline{\textbf{return}}$ $M''\cup \{e\in M'|e\cap \bigcup_{e'\in M''}e'=\emptyset\}$;\\
\noalign{\smallskip} \hline
\end{tabular}
\end{small}
\caption{\textsc{RM} algorithm.}
\end{center}
\end{figure}
\cite{Talman}provides two dynamic procedures that produce constrained Walrasian equilibrium.
But it does not address the computation issues, and the third condition of constrained Walrasian equilibrium cannot be guaranteed either. In order to make sure that all the items whose prices exceed
their lower bound prices will be sold(the third criterion of constrained Walrasian equilibrium),
the $\textsc{RM}$
subroutine shown in Figure 2 is called in step 9. Given a demand
situation $\mathcal{D}=(D_i)_{i\in N}$, a partial matching $M$
consistent with $\mathcal{D}$, the current price vector
$\textbf{p}$, and the lower bound price vector
$\underline{\textbf{p}}$, $\textsc{RM}$ returns a matching $M'$
such that \textbf{(1)} $\pi^{M\cup M'}$ is an equilibrium
allocation, \textbf{(2)} $M\cap M'=\emptyset$, and \textbf{(3)}
$\{a\in X\setminus\bigcup_{e\in
M}e|\textbf{p}_a>\underline{\textbf{p}}_a\}\subseteq\bigcup_{e\in
M'}e$.
Observe MAPR and RM subroutine. We can find that:
\begin{itemize}
\item computation of each step is polynomial in $|N|$ and $|X|$;
\item for each $t\geq 0$, the number of the loops consisting of
steps 3-5 is not more than $|X|$; and
\item the number of the loops consisting of steps 2-8 is not more
than $\sum\limits_{a\in
X}(\overline{\textbf{p}}_a-\underline{\textbf{p}}_a)$.
\end{itemize}
Consequently, MAPR always terminates and is polynomial in $|N|$,
$|X|$, and $\sum\limits_{a\in
X}(\overline{\textbf{p}}_a-\underline{\textbf{p}}_a)$.
In order to prove the correctness of MAPR and RM, we will first
give some definitions and provide three lemmas, then we will prove
that MAPR can lead to a constrained Walrasian equilibrium with the
help of these three lemmas. In the following discussion, we
suppose that MAPR terminates at some time $T\geq 0$;
$\textbf{p}^t$, $M^t$, $R^t$ ($R^t(i,a)=R_i[a]$ for all $\langle
i,a\rangle\in N\times X$, where $R_i$ is the vector kept by buyer
$i$ at time $t$), and $(D_i^t)_{i\in N}$ denote the price vector,
partial matching that has been made so far, rationing system, and
demand situation at time $0\leq t\leq T$, respectively. Let
$X^t=\{a\in X\setminus\bigcup_{e\in
M^t}e|\textbf{p}^t_a>\underline{\textbf{p}}_a\}$ and $N^t=\{i\in
N\setminus\bigcup_{e\in M^t}e| D^t_i\cap X^t\neq\emptyset\}$.
Now we introduce three auxiliary lemmas (in which
$\mathcal{D}=(D_i)_{i\in N}$ denotes a demand situation). These
three lemmas are closely connected. The proof of Lemma 4 is based
on Lemma 2 and Lemma 3, and the proof of Theorem 2 is based on the
these three lemmas. Lemma 2 states that, each nonempty subset of a
minimal over-demanded set of items is not under-demanded.
\begin{Lemma}
Let $X'$ be a minimal over-demanded set of items. Then for
each $\emptyset\subset X''\subseteq X'$, $|\{i\in N|D_i\cap
X''\neq\emptyset\textrm{ and }D_i\subseteq X'\}|>|X''|$.
\end{Lemma}
The proof of Lemma 2 is not very hard, and comes from using the
reduction to absurdity.
Lemma 3 states that, the cardinality of a maximum matching is not
less than the cardinality of a set of real items if each
subset of the set is not under-demanded.
\begin{Lemma}
Let $X'\subseteq X\setminus\{o\}$ and $|\{i\in N|D_i\cap
X''\neq\emptyset\}|\geq|X''|$ for each $X''\subseteq X'$. If $M$
is a maximum matching of $BG((D_i\setminus\{o\})_{i\in N})$, then
$|M|\geq|X'|$.
\end{Lemma}
The proof of Lemma 3 is similar to that of Theorem 1. Due to lack
of space, it is omitted.
Lemma 4 states that, all the items in $X^t$ can be sold. The proof of Lemma 4 is based on Lemma 2 and Lemma 3.
\begin{Lemma}
Let $\mathcal{D}^t=(D^t_i\cap X^t)_{i\in N^t}$. Then
$|\hat{M}_{\mathcal{D}^t}|=|X^t|$ for each $0\leq t\leq T$.
\end{Lemma}
\textsc{Proof.} \begin{small}We first prove that $|\{i\in
N^t|D^t_i\cap X'\neq\emptyset\}|\geq|X'|$ for each
$\emptyset\subset X'\subseteq X^t$ and $0\leq t\leq T$:
\begin{enumerate}
\item It holds at $t=0$ because $X^0=\emptyset$.
\item Suppose MAPR does not stop at $\hat{t}\geq 0$ and $|\{i\in
N^t|D^t_i\cap X'\neq\emptyset\}|\geq|X'|$ for each
$\emptyset\subset X'\subseteq X^t$ and $0\leq t\leq \hat{t}$.
\item Then $X_{min}\neq\emptyset$ and $\overline{X}$ are computed
at time $\hat{t}$ and steps 6-7 of MAPR. Pick any
$\emptyset\subset X'\subseteq X^{\hat{t}+1}$. Let $N_1=\{i\in
N^{\hat{t}}|D_i^{\hat{t}}\subseteq X_{min}\textrm{ and
}D_i^{\hat{t}}\cap X'\neq\emptyset\}$ and $N_2=\{i\in
N^{\hat{t}}|D_i^{\hat{t}}\cap (X'\setminus
X_{min})\neq\emptyset\}$. There are two possibilities:
\begin{itemize}
\item[Case I]: $\overline{X}=\emptyset$. So
$X^{\hat{t}+1}=X^{\hat{t}}\cup X_{min}$. According to Lemma 2 and
item 2, we have $|N_1|>|X'\cap X_{min}|$ and
$|N_2|\geq|X'\setminus X_{min}|$. It is easy to find that
$D_i^{\hat{t}+1}\cap X'\neq\emptyset$ for each $i\in N_1\cup
N_2\subseteq N^{\hat{t}+1}$ and $N_1\cap N_2=\emptyset$. So
$|\{i\in N^{\hat{t}+1}|D^{\hat{t}+1}_i\cap
X'\neq\emptyset\}|\geq|N_1\cup N_2|=|N_1|+|N_2|>|X'\cap
X_{min}|+|X'\setminus X_{min}|=|X'|$.
\item[Case II]: $\overline{X}\neq\emptyset$ and some $a\in
\overline{X}$ is assigned to some buyer $j$ such that $a\in
D_j^{\hat{t}}\subseteq X_{min}$. So
$X^{\hat{t}+1}=X^{\hat{t}}\setminus\{a\}$. According to Lemma 2
and item 2, we have $|N_1|>|X'\cap X_{min}|$ and
$|N_2|\geq|X'\setminus X_{min}|$. It is easy to find that
$D_i^{\hat{t}+1}\cap X'\neq\emptyset$ for each $i\in
(N_1\setminus\{j\})\cup N_2\subseteq N^{\hat{t}+1}$ and $N_1\cap
N_2=\emptyset$. Consequently, $|\{i\in
N^{\hat{t}+1}|D^{\hat{t}+1}_i\cap
X'\neq\emptyset\}|\geq|(N_1\setminus\{j\})\cup
N_2|\geq|N_1|-1+|N_2|\geq|X'\cap X_{min}|+|X'\setminus
X_{min}|=|X'|$.
\end{itemize}
Consequently, $|\{i\in N^{\hat{t}+1}|D^{\hat{t}+1}_i\cap
X'\neq\emptyset\}|\geq|X'|$.
\end{enumerate}
According to items 1--3, $|\{i\in N^t|D^t_i\cap
X'\neq\emptyset\}|\geq|X'|$ for each $X'\subseteq X^t$ and $0\leq
t\leq T$. It is easy to find that
$|\hat{M}_{\mathcal{D}^t}|\leq|X^t|$ for each $0\leq t\leq T$.
According to Lemma 3, we have $|\hat{M}_{\mathcal{D}^t}|\geq|X^t|$
for each $0\leq t\leq T$. So $|\hat{M}_{\mathcal{D}^t}|=|X^t|$ for
each $0\leq t\leq T$.$\quad\square$
\end{small}
Now we are ready to establish the following correctness theorem
for MAPR (and RM subroutine).
\begin{Theorem}
$\langle\textbf{p}^T,R^T,\pi^{M^T}\rangle$ found by MAPR, is a
constrained Walrasian equilibrium.
\end{Theorem}
\textsc{Proof.} (Sketch)
\begin{small}$\langle\textbf{p}^T,R^T,\pi^{M^T}\rangle$ is a
constrained Walrasian equilibrium iff it satisfies the five
conditions shown in page 2.
\begin{enumerate}
\item It is easy to find that conditions (1), (4), and (5) are
satisfied by $\langle\textbf{p}^T,R^T,\pi^{M^T}\rangle$.
\item For each buyer $i$ and the item assigned to her
$a=\pi^{M^T}(i)$, there are two possibilities: Case I (step (8)),
$i$ is the winner of a lottery on item $a$ at some time
$T'\leq T$, and Case II (step (6) and (9)), $a$ is assigned to $i$
at time $T$.
\begin{enumerate}
\item In case I, $a\in D_i(\textbf{p}^{T'},R^{T'})$. So
$u_i(a)-\textbf{p}^{T'}_a\geq u_i(b)-\textbf{p}^{T'}_b$ for all
$b\in \{b\in X|R^{T'}(i,b)=1\}$. Because
$R^{T'}(i,a)=R^{T}(i,a)=1$, $\textbf{p}^{T'}_a=\textbf{p}^{T}_a$,
$R^{T'}(i,b)\geq R^{T}(i,b)$ and
$\textbf{p}^{T'}_b\leq\textbf{p}^{T}_b$ for all $b\in X$,
$u_i(a)-\textbf{p}^{T}_a\geq u_i(b)-\textbf{p}^{T}_b$ for all
$b\in\{b\in X|R^{T}(i,b)=1\}$. So $a\in
D_i(\textbf{p}^{T},R^{T})$.
\item In case II, according to the definition of $\pi^{M^T}$ (see
RM subroutine and steps (6)--(9)), we have $a\in
D_i(\textbf{p}^{T},R^{T})$.
\end{enumerate}
Consequently, $\pi^{M^T}$ is an equilibrium allocation.
\item According to Lemma 4, all the items in $X^T$ are sold.
Consequently, $\textbf{p}^T_a=\underline{\textbf{p}}_a$ for each
$a\in\{b\in X|(\forall i\in N)\pi^{M^T}(i)\neq b\}$. The
correctness of RM subroutine can derive from item 2 and item 3
directly.
\end{enumerate}
So $\langle\textbf{p}^T,R^T,\pi^{M^T}\rangle$ is a constrained
Walrasian equilibrium.$\quad\square$
\end{small}
\begin{Example}
\begin{small}
See Example 1. Apply MAPR to $\langle
E,\underline{\textbf{p}},\overline{\textbf{p}}\rangle$. The
demands, price vectors, rationing system and other relevant data
generated by MAPR are illustrated in Table 2, where $U_i$, $D_i$,
$X'$, $N'$, and $X_{min}$ denote $\{a\in X|R^t(i,a)=0\}$,
$D_i(\textbf{p}^t,R^t)$, $X\cap\bigcup_{e\in M^t}e$,
$N\cap\bigcup_{e\in M^t}e$, and the value of $X_{min}$ computed by
the seller at step (6) and time $t$.
At $t=3$, the price of $c$ has reached its upper bound 4. The
seller assigns randomly $c$ to buyer $2$ or buyer $3$. So there
are two different possible histories of resource allocation from
$t=3$. Along the history of $t=4.1;5.1;6.1$, MAPR finds
$\langle\textbf{p}^{6.1},R^{6.1},\pi^{M^{6.1}}\rangle$, where
$\pi^{M^{6.1}}(1)=o$, $\pi^{M^{6.1}}(2)=c$, $\pi^{M^{6.1}}(3)=b$,
$\pi^{M^{6.1}}(4)=a$, and $\pi^{M^{6.1}}(5)=d$. Along the history
of $t=4.2;5.2;6.2$, MAPR finds
$\langle\textbf{p}^{6.2},R^{6.2},\pi^{M^{6.2}}\rangle$, where
$\pi^{M^{6.2}}(1)=o$, $\pi^{M^{6.2}}(2)=b$, $\pi^{M^{6.2}}(3)=c$,
$\pi^{M^{6.2}}(4)=a$, and $\pi^{M^{6.2}}(5)=d$.
\end{small}
\end{Example}
\begin{table*}
\caption{Data Generated by MAPR}
\begin{center}
\begin{small}
\begin{tabular}{c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c}
\hline
t & $\textbf{p}^t_o$ & $\textbf{p}^t_a$ & $\textbf{p}^t_b$ & $\textbf{p}^t_c$ & $\textbf{p}^t_d$ & $X_{min}$ & $U_1$ & $U_2$ & $U_3$ & $U_4$ & $U_5$ & $N'$ & $D_1$ & $D_2$ & $D_3$ & $D_4$ & $D_5$ & $X'$\\
\hline
0 & 0 & 5 & 4 & 1 & 5 & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\{c\}$ & $\{c\}$ & $\{c\}$ & $\{a\}$ & $\{d\}$ & $\emptyset$\\
1 & 0 & 5 & 4 & 2 & 5 & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\{c\}$ & $\{c\}$ & $\{c\}$ & $\{a\}$ & $\{d\}$ & $\emptyset$\\
2 & 0 & 5 & 4 & 3 & 5 & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\{c,d\}$ & $\{c\}$ & $\{c\}$ & $\{a\}$ & $\{d\}$ & $\emptyset$\\
3 & 0 & 5 & 4 & 4 & 5 & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\{d\}$ & $\{c\}$ & $\{c\}$ & $\{a\}$ & $\{d\}$ & $\emptyset$\\
4.1 & 0 & 5 & 4 & 4 & 5 & $\{d\}$ & $\emptyset$ & $\emptyset$ & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\{2\}$ & $\{d\}$ & & $\{d\}$ & $\{a\}$ & $\{d\}$ & $\{c\}$\\
5.1 & 0 & 5 & 4 & 4 & 6 & $\{d\}$ & $\{c\}$ & $\emptyset$ & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\{2\}$ & $\{d\}$ & & $\{b,d\}$ & $\{a\}$ & $\{d\}$ & $\{c\}$\\
6.1 & 0 & 5 & 4 & 4 & 7 & $\emptyset$ & $\{c\}$ & $\emptyset$ & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\{2\}$ & $\{o,d\}$ & & $\{b\}$ & $\{a\}$ & $\{d\}$ & $\{c\}$\\
4.2 & 0 & 5 & 4 & 4 & 5 & $\{d\}$ & $\emptyset$ & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\{3\}$ & $\{d\}$ & $\{a,b\}$ & & $\{a\}$ & $\{d\}$ & $\{c\}$\\
5.2 & 0 & 5 & 4 & 4 & 6 & $\{d\}$ & $\{c\}$ & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\{3\}$ & $\{d\}$ & $\{a,b\}$ & & $\{a\}$ & $\{d\}$ & $\{c\}$\\
6.2 & 0 & 5 & 4 & 4 & 7 & $\emptyset$ & $\{c\}$ & $\{c\}$ & $\emptyset$ & $\emptyset$ & $\emptyset$ & $\{3\}$ & $\{o,d\}$ & $\{a,b\}$ & & $\{a\}$ & $\{d\}$ & $\{c\}$\\
\hline
\end{tabular}
\end{small}
\end{center}
\end{table*}
\section{Expected profits, Expected Prices, and Strategical Issues}
Since the history of MAPR is nondeterministic, we need to
introduce concepts of buyers' \emph{expected profits} and
items' \emph{expected prices}. Let $R_*^t$ be a rationing
system s.t. $R_*^t(i,a)=1$ if $\{i,a\}\in M^t$ or
$a\not\in\bigcup_{e\in M^t}e$, and 0 otherwise. Because we can
induce $M^t$ from $R_*^t$. So $M^t$ can be written as $M^{R_*^t}$.
We say $\langle \textbf{p}^t,R_*^t\rangle$ is an allocation
situation. Assume that the computation of $\textsc{MODS}$
algorithm and the selection of items in step (8) are
deterministic, all the lots happening in MAPR are fair
\footnote{Suppose there are $k$ buyers drawing lots for the right
to buy item $a$. Then the lot is fair if each one of these
buyers has $1/k$ chance of winning the lot.}. Then buyer $i$'s
expected profit and item $a$'s expected price on $\langle
\textbf{p},R\rangle$ (i.e., $u^*_i(\textbf{p},R)$ and
$\textbf{p}^*_a(\textbf{p},R)$) are:
\begin{small}
\begin{displaymath}
u^*_i(\textbf{p},R)=\left\{\begin{array}{ll} V_{i}(\textbf{p},R) & \textrm{if}\:X_{min}=\emptyset\\
u^*_{i}(\textbf{p}',R) &
\textrm{if}\:\overline{X}=\emptyset\\
\frac{\sum_{i'\in N'}u^*_{i}(\textbf{p},R_{i'})}{|N'|} &
\textrm{otherwise}
\end{array} \right.
\end{displaymath}
\begin{displaymath}
\textbf{p}_a^*(\textbf{p},R)=\left\{\begin{array}{ll} \textbf{p}_a & \textrm{if}\:X_{min}=\emptyset\\
\textbf{p}_a^*(\textbf{p}',R) &
\textrm{if}\:\overline{X}=\emptyset\\
\frac{\sum_{i'\in N'}\textbf{p}_a^*(\textbf{p},R_{i'})}{|N'|} &
\textrm{otherwise}
\end{array} \right.
\end{displaymath}
where (let $\mathcal{D}=(D_i(\textbf{p},R))_{i\in N}$):
\begin{itemize} \item $X_{min}=\emptyset$ if $|\hat{M}_{\mathcal{D}}|=|\{i\in N|o\not\in D_i(\textbf{p},R)\}|$, and $\textsc{MODS}(\mathcal{D},\hat{M}_{\mathcal{D}})$
otherwise; $\overline{X}=\{a\in
X_{min}|\textbf{p}_a=\overline{\textbf{p}}_a\}$;
\item $\textbf{p}'_{a}=\textbf{p}_{a}$ for all $a\not\in X_{min}$
and $\textbf{p}'_{a}=\textbf{p}_{a}+1$ for all $a\in X_{min}$;
\item $b\in \overline{X}$ is the item selected by the
seller in step (8);
\item $N'=\{i\in N|b\in D_i(\textbf{p},R)\subseteq X_{min}\}$;
\item for all $\langle i,a\rangle\in N\times X$:
$R_{i'}(i,a)=R(i,a)$ if $a\neq b$; $R_{i'}(i,b)=0$ if $i\neq i'$;
and $R_{i'}(i',b)=1$.
\end{itemize}
\end{small}
In fact, $u^*_i(\textbf{p},R)$ and $\textbf{p}^*_a(\textbf{p},R)$
can be computed by developing a search tree: each node is an
allocation situation, and is expanded (if $X_{min}\neq\emptyset$)
into (i) one single branch if $\overline{X}=\emptyset$, and (ii)
$|N'|$ branches otherwise. See Table 1 and Table 2. We can find
that
$u^*_1(\textbf{p}^0,R^0_*)=0.5*u^*_1(\textbf{p}^{6.1},R^{6.1}_*)+0.5*u^*_1(\textbf{p}^{6.2},R^{6.2}_*)=0$,
$u^*_3(\textbf{p}^0,R^0_*)=0.5*u^*_3(\textbf{p}^{6.1},R^{6.1}_*)+0.5*u^*_3(\textbf{p}^{6.2},R^{6.2}_*)=2.5$,
$\textbf{p}_a^*(\textbf{p}^0,R^0_*)=0.5*\textbf{p}_a^*(\textbf{p}^{6.1},R^{6.1}_*)+0.5*\textbf{p}_a^*(\textbf{p}^{6.2},R^{6.2}_*)=5$.
As most collective decision mechanisms, MAPR is generally not
\emph{strategyproof} (in the sense of expected profit). For
instance, see Example 4. If buyer 1 reports her demands sincerely,
then her expected profit is 0. However, if 1 knows other buyers'
valuations and reports strategically, then she reports $\{c\}$
from $t=0$ to $t=3$ (i.e., as if her valuation to item $c$ is
not less than 7), then reports sincerely, then her expected
profit changes to 1/3, which makes her better off.
Now we are interested in two questions: (1) is MAPR strategyproof
for some restricted domains? (2) when it is not, how hard is it
for an buyer who knows the valuations of the others to compute an
optimal strategy?
First we define reporting strategies and manipulation problems
formally. Without loss of generality, let 1 be the manipulator.
Note that not every sequence of 1's demands is reasonable. For
instance, see Example 4 and Table 2. The seller can detect 1's
manipulation if 1 reports $\{c\}$, $\{c\}$, $\{c,d\}$, and $\{c\}$
at $t=0$, 1, 2, and 3, respectively, because there is no value
function $u$ s.t.
$u(c)-\textbf{p}^2_c=u(c)-3=u(d)-5=u(d)-\textbf{p}^2_d=u(d)-\textbf{p}^3_d<u(c)-\textbf{p}^3_c=u(c)-4$.
A \emph{strategy} for buyer 1 is a value function $u: X
\rightarrow \mathbb{Z}_+$ with $u(o)=0$. So 1 can safely
manipulate the process of MAPR when she reports her demands
according to $u$ completely (as if $u$ is her true value
function). A \emph{manipulation problem} M (for buyer 1) is a
5-tuple $\langle N,X,\{u_i\}_{i\in N}, \underline{\textbf{p}},
\overline{\textbf{p}}\rangle$ where $\langle N,X,\{u_i\}_{i\in
N}\rangle$ is an economy, $\underline{\textbf{p}}$ and
$\overline{\textbf{p}}$ are the lower and upper bound price
vectors on $X$, respectively. A strategy for M is \emph{optimal}
if 1 can not strictly increase her expected profit by reporting
her demands according to any other strategy.
Now, back to question (1): we show that the answer is positive
when there are two buyers.
\begin{Theorem}
Let $M=\langle N,X,\{u_i\}_{i\in N}, \underline{\textbf{p}},
\overline{\textbf{p}}\rangle$ be a manipulation problem s.t.
$N=\{1,2\}$. Then $u_1$ is optimal for $M$.
\end{Theorem}
\textsc{Proof.}
\begin{small}
Suppose that if 1 reports sincerely, then her expected profit is
$\Delta$. Let $D_1$ and $D_2$ be 1 and 2's true demands at
$\underline{\textbf{p}}$ and $R$ respectively, where $R(i,a)=1$
for each $i\in N$ and $a\in X$.
Obviously, if $D_1\cup D_2=\{o\}$ or $|D_1\cup D_2|\geq 2$ (i.e.,
$X_{min}=\emptyset$ at $t=0$) then $\Delta=\max_{a\in
X}(u_1(a)-\underline{\textbf{p}}_a)$, which is the best possible
outcome for 1. So $u_1$ is optimal in these cases.
Now, suppose $D_1=D_2=\{a\}$ s.t. $a\neq o$. Pick any strategy
$u'$. Let $k=\overline{\textbf{p}}_a-\underline{\textbf{p}}_a$,
$k_i=u_i(a)-\underline{\textbf{p}}_a-\max_{b\in
X\setminus\{a\}}(u_i(b)-\underline{\textbf{p}}_b)$, $b_i\in
X\setminus\{a\}$ s.t. $u_i(b_i)-\underline{\textbf{p}}_{b_i}
=u_i(a)-\underline{\textbf{p}}_a-k_i$, and
$\hat{k}=\min(k,k_1-1,k_2-1)$. Then if 1 applies strategy $u_1$,
then she will report $D_1$ from $t=0$ to $t=\hat{k}$ and:
\begin{enumerate}
\item if $\hat{k}=k$, then $\Delta=0.5\ast
(u_1(a)-\underline{\textbf{p}}_a-k)+0.5\ast(u_1(b_1)-\underline{\textbf{p}}_{b_1})$=$u_1(b_1)-\underline{\textbf{p}}_{b_1}+0.5\ast(k_1-k)>u_1(b_1)-\underline{\textbf{p}}_{b_1}$.
If 1 applies $u'$ instead, then her expected profit will not be
better than $u_1(b_1)-\underline{\textbf{p}}_{b_1}<\Delta$ if $
u'(a)-\underline{\textbf{p}}_a-\max_{b\in
X\setminus\{a\}}(u'(b)-\underline{\textbf{p}}_b)\leq k$, and will
not be better than $\Delta$ otherwise.
\item if $k>\hat{k}=k_1-1$, then
$\Delta=u_1(b_1)-\underline{\textbf{p}}_{b_1}$. Because 2 can
insist on $\{a\}$ to $t=\min(k,k_2-1)\geq k_1-1$, 1's expected
profit can not be better than $\Delta$.
\item if $k>\hat{k}=k_2-1$, then
$\Delta=u_1(a)-\underline{\textbf{p}}_{a}-k_2\geq
u_1(a)-\underline{\textbf{p}}_{a}-k_1=u_1(b_1)-\underline{\textbf{p}}_{b_1}$.
Because 2 can insist on $\{a\}$ to $t=k_2-1$, 1's expected profit
can not be better than $\Delta$.
\end{enumerate}
To sum up, in all cases, 1 can not strictly increase her expected
profit by applying strategy $u'$. So $u_1$ is optimal for $M$.
$\quad\square$
\end{small}
For the cases where there are more than two buyers, we conjecture
that the manipulation problem is NP-hard, but we could not find a
proof.
\section{Conclusion}
We have presented a decentralized protocol for allocating
indivisible resources under price rigidities, and proved formally
that it can discover constrained Walrasian equilibria in
polynomial time. We also have studied the protocol from the points
of computation of buyers' expected profits and items'
expected prices, and discussed the manipulation (by one buyer)
problem in the sense of buyer's expected profit. There are
several directions for future work. One direction would be to
prove the conjecture about the complexity of manipulation (in the
sense of expected profits) by one buyer. Another direction would
be to study manipulation (in the sense of expected prices) by one or
more buyers (whose manipulation motivation is not to buy some
resources but to put up the prices of some resources).
Furthermore, we plan to study the problems of allocating divisible
resources \cite{Brams5} and sharable resources \cite{Airiau} under
prices rigidities.
\section*{Acknowledgments}
This work is supported by the National Natural Science Foundation
of China under grant No.61070232, No.61105039, No.61272295, and
the Fundamental Research Funds for the Central Universities (Issue
Number WK0110000026).
\bibliographystyle{named}
|
\section{Introduction}
\label{sec:intro}
Accreting X-ray pulsars \citep[see, e.g.,][for a recent
review]{caballero:12a} exhibit complex and highly energy-dependent
pulse profiles \citep{lutovinov:09a}. These profiles can vary with
time and luminosity, for instance over outbursts of transient sources
\citep[e.g.,][]{tsygankov:06a} or in the case of Her X-1 with a
super-orbital period \citep{staubert:13a}. However, the detailed
physical origin for this complexity is still poorly understood. A
number of efforts in the past have attempted to disentangle the components of intrinsic
beam patterns for individudal sources by a complex ``backward''
approach as described by \citet{kraus:95a}, by Fourier-decomposing the
observed pulse profiles into individual contributions of radiation
from the poles \citep[e.g.,][]{kraus:96a, caballero:11a, sasaki:12a}.
While this approach has proven to be very successful in offering
possible solutions for the beam patterns, that is, the angle- and
energy-dependent emissivity of X-rays at each pole, it does not
address the physical processes that yield these beam patterns, nor does
it result in a unique choice among several possible solutions.
The most striking feature of many observations is that X-ray pulse
profiles are strongly energy dependent. This is especially apparent in
the energy-dependent location of the main pulse peak, which changes
significantly in pulse phase, that is, the main peak shows a phase lag
\citep{tsygankov:07a,tsygankov:06a,ferrigno:11a}. This phase lag, relative to the mean pulse profile, is
most apparent around the energy of the cyclotron line, at
energies where photons can scatter resonantly with electrons on Landau
orbits in the strong magnetic field of the accretion column and where
the scattering cross section is strongly angle dependent.
In this \textsl{Letter}, we present a new approach in which pulse
profiles are calculated based on a self-consistent physical model of
the intrinsic radiation pattern at the two magnetic poles, including
all general relativistic effects such as light bending or
gravitational redshifting (Sect.~\ref{model}). This approach follows
early fundamental works on the radiation from cyclotron line sources
\citep[see, e.g.,][and references therein]{meszaros:88b,sturner:94a,zheleznyakov:86a, soffel:85a}, but differs
in numerical method. This
allows for the resolution of model spectra and pulse profiles at a new level, as we
illustrate here for one characteristic observable, energy-dependent
pulse profiles (Sect.~\ref{pp_maps}). We explore under which
conditions intrinsic deviations from the mean, energy averaged pulse profile due to cyclotron resonance scattering lead to the formation of
characteristic phase lags relative to the mean pulse
profile at the cyclotron energies.
\section{Model setup}
\label{model}
We considered a canonical neutron star ($M=1.4\,M_{\odot}$, $R=10$\,km)
with a strong magnetic field on the order of several $10^{12}$\,G
that accretes matter onto two magnetic poles (see Fig.~\ref{fig:bp}). We furthermore assumed that we observe only lateral emission from the walls of possibly elongated accretion columns, a ``fan beam scenario''. For sources with clearly defined cyclotron lines, this necessarily induces the postulation of a constrained
``efficient area of the accretion column to line formation'' \citep{nishimura:08a}, which must be of practicable negligible magnetic field gradient.
The X-ray emitting region of the accretion column at each pole is
therefore modeled by a simple, homogeneous, small cylindrical volume. The location, radius, and
height of the cylindrical volume, the location of the accretion columns
with respect to the spin axis of the neutron star, and the inclination
of the system with respect to the observer are free model parameters. We do not address any possible ``pencil beam'' radiation components, nor do we consider surface reflection \citep[see, e.g.,][]{poutanen:13a} for this pilot study.
The plasma electrons have a Maxwellian temperature distribution
parallel to the $B$-field and are magnetically quantized in their
perpendicular momenta. We generated the intrinsic beam pattern at each
pole by reprocessing X-ray continuum seed photons from the column by
cyclotron resonant scattering between electrons and X-ray photons in
its outer layer. This picture is motivated by the comparably very
small mean free path of resonant photons that translates into a thin
``scattering atmosphere'' around the optically thick core. The
scattering of photons and electrons causes cyclotron resonant
scattering features (if observable) in the source spectrum at $E_{n,
\mathrm{cyc}}\sim n \, 11.6\,B\,{[10^{12}}\,$G$]/
(1+z)\,\mathrm{keV}$, where $z$ is the gravitational redshift and $n$
is an integer. In the following we assume $B= 4 \times
10^{12}$\,G, a parallel electron temperature of
$kT_\mathrm{e}=3$\,keV, and a cyclotron scattering atmosphere with
Thomson optical depth $\tau =10^{-3}$ perpendicular to the $B$-field
axis \citep[see, e.g.,][]{schoenherr:07a, suchy:08a}.
\begin{figure}
\centering
\resizebox{0.8\hsize}{!}{\includegraphics{figures/figure1}}
\caption{Observer's view of a neutron star, illustrating the general
relativistic effects on the visibility and flux of two accretion
columns for an observer at infinite distance from the neutron
star. Sky coordinates are parameterized through the impact
parameters $A$ and $B$, which correspond to the projected distance
from the gravitational center in units of the Schwarzschild
radius, $r_\mathrm{s}$. The illustration corresponds to the
two-pole geometry used in Sect.~\ref{sect:asym}. Different
apparent sizes of the accretion columns are due to strong light-bending effects.}
\label{fig:bp}
\end{figure}
\begin{figure*}\sidecaption
\centering
\includegraphics[width=5.5cm]{figures/{figure2a}.pdf}
\includegraphics[width=5.5cm]{figures/{figure2b}.pdf}
\caption{ \label{fig:bp0} Energy-resolved beam patterns after
cyclotron scattering for a cylindrical geometry (``fan
beam''). \textbf{Left:} case of a static plasma.
\textbf{Right:} plasma with bulk velocity $v=0.4c$. The
black line shows the angular distribution of all photons.
The red, green, and blue lines mark photons escaping at
energies close to the cyclotron resonances for the
respective angles (not to scale). While the photons
that escape at the fundamental resonance (red) have all
undergone scattering, the photons escaping at the higher
resonances are unaltered from the continuum
\citep[see][]{schwarm:13a}.}
\end{figure*}
Because we wish to investigate the isolated effects of beaming by cyclotron
resonant scattering on the observable pulse profiles separated from
possible continuum effects, we assumed the continuum radiation to be
isotropic and to have a simple power-law spectrum with a photon index
$\Gamma = 1$. The assumption of isotropy is justified as the
intrinsic beaming of the continuum radiation is expected to be
comparably weak. For a fast falling plasma, an overall downward
beaming has to be taken into account as well
\citep[see][]{lyubarsky:88a}. The energy and angular redistribution
of the continuum while passing the cyclotron scattering layer were
calculated based on Monte Carlo simulations of
\citet[][2013]{schwarm:10a}, which follow the same fundamental
physical setup as described in detail by \citet[][]{araya:99a}, \citet[][]{schoenherr:07a}, and references therein. Fully relativistic cross-sections \citep{sina:96a} were used to describe the
polarization-averaged scattering process. Again, to bring out the
effects of cyclotron scattering on the pulse profiles, identical
optical depths and $B$-field strengths were assumed for both poles. We
obtain the flux emitted from each point of the surface of the walls of
the cylindrical volume as a function of energy and angle with respect
to the $B$-field.
We considered both static volumes and a falling plasma with a bulk
velocity of $0.4c$ (see Fig.~\ref{fig:bp0}). The effect of a bulk
velocity on our simulation results is discussed in more detail
by \citet{schwarm:13a}. For earlier work on cyclotron
resonant scattering in non-static plasmas see, e.g.,
\citet{soffel:85a}, \citet{isenberg:98a}, and \citet{serber:00a}.
As the large-scale $B$-field structure of High-Mass X-ray Binaries and the magnetospheric
coupling of the accretion flow are poorly constrained, we considered
both a symmetric geometry and a geometry where the magnetic poles are
displaced in longitude and latitude with respect to an antipodal
setting. The latter has been shown to be one option to generate
asymmetric pulse profiles that resemble observed ones from otherwise
symmetric assumptions \citep[e.g.,][]{bulik:95a, kraus:96a,
caballero:11a, sasaki:12a}. We calculated the observed time and
energy-dependent flux from these setups with a fully relativistic
ray-tracing code that includes light bending, gravitational
redshifting, and special relativistic effects due to the motion of the
emitter \citep{falkner:13b}.
\section{Energy-dependent pulse profiles}\label{pp_maps}
\subsection{Symmetric geometry}
\label{sect:sym}
First, we discuss a symmetric, antipodal geometry of two static
accretion columns of radius $r_1=r_2=1$\,km and height
$h_1=h_2=100$\,m, which are displaced at an angle $i_1 = i_2 =
35^\circ$ with respect to the spin axis of the neutron star. Because
we assumed a constant magnetic $B$-field, the heights $h_i$ do not
represent a realistic accretion column structure. They instead rather
reflect the uncertainty of the height from which emission from a
sufficiently constrained line-forming region is observed, which for
this example was assumed to be somewhere at the base of the column
not farther than $100\,$m from the surface.
Figure~\ref{fig:maps_obs} shows the resulting energy-dependent pulse
profiles for different observing angles. Horizontal cuts through the
figures represent pulse profiles at a given energy, while vertical
cuts represent the deviation of the X-ray spectrum from a normalizing
continuum at a given pulse phase. Observing this system at a line of
sight tilted by $i_\mathrm{obs}=25^\circ$ with respect to the
rotational axis results in practically no features being apparent from which one might distinguish between the relative pulse profiles at different
energies. Observations would show one symmetric broad pulse over the
entire energy range. With increasing $i_\mathrm{obs}$, the main peak
of the pulse profile becomes narrower and strong deviations of the
pulse profiles appear around the redshifted cyclotron line energies
($\sim$32, 64, and 96\,keV). These changes can be understood from the
very strong perpendicular beaming of photons that escape at the
cyclotron energies (Fig.~\ref{fig:bp0}). These photons can come into
view of the observer, for a suitable geometric setup, at a pulse phase
when most of the main beam at all other energies is still hidden.
\begin{figure*}\sidecaption
\centering
\includegraphics[width=12.5cm]{figures/figure3}
\caption{\label{fig:maps_obs}Energy-dependent flux as a
function of pulse phase for different observation angles
$i_\mathrm{obs}=25^\circ$, $50^\circ$, and $75^\circ$ from left to
right (see Sect.~\ref{sect:sym} for the geometrical setup). The pulse
profiles are normalized such that for every energy bin the mean flux
is zero and the corresponding standard
deviation is unity \citep[after][]{ferrigno:11a}.
The pulse profiles are plotted twice for clarity.}
\end{figure*}
\subsection{``Wavy'' phase lags}
\label{sect:asym}
Motivated by an observational study of phase lags, we demonstrate the
effects of strong asymmetry of the location of the accretion columns
and of a non-static plasma for one specific example: To explain the
strong changes in the highly asymmetric and variable pulse profiles of
the accreting X-ray pulsar \object{4U0115+63}, \citet{ferrigno:11a}
invoked a two-pole geometry with $i_1 =74^\circ$, $i_2=32^\circ$ and a
phase offset of $68^\circ$ for an observation angle of
$i_\mathrm{obs}=60^\circ$, as was proposed by
\citet{sasaki:12a} from applying the pulse decomposition method to
this source. The authors additionally assumed laterally emitting
columns of $r_1=r_2 \sim 700$\,m and $h_1=h_2=2$\,km
(Fig.~\ref{fig:bp}) and heuristically modeled the intrinsic emissivity
pattern from these columns as the sum of an upward and a downward
pointing beam with a Gaussian emissivity profile. From purely
geometric considerations they proposed that a relatively enhanced
upward component at the cyclotron line energies might be causing the
observed pulse profile changes and the associated strong phase shifts.
We now adopt the same parameters for the two-pole accretion column
geometry of our numerical model. The relative fluxes generated
from a constrained emitting volume of given parameters -- in
particular a given value for the magnetic field $B$ for a source
with sharp cyclotron lines -- located at arbitrary height between
the top of the accretion mound and $2\,$km up in the column above
the surface show very little variation. Therefore, we used spatially
averaged fluxes (Fig.~\ref{fig:wave}) for our further analysis.
In contrast to \citet{ferrigno:11a} we did not assume any artificially
imposed beamed emission patterns. Instead, as discussed for the
symmetric setting in the previous paragraph, we used the simulated
intrinsic beam patterns of the photons that were redistributed
in angle by cyclotron resonant scattering \citep[see
Fig.~\ref{fig:bp0} and][]{schoenherr:07a,schwarm:13a}.
Figure~\ref{fig:wave} shows the energy-dependent pulse profiles and
associated phase lags for a static plasma and for a non-static plasma
with a bulk velocity of $v=0.4c$ within the line-forming region.
While for the static plasma the phase shifts of the peak at the
cyclotron energies relative to the mean pulse profile are very
localized and rather small, the introduction of a bulk velocity
component yields a broad and wavy morphology similar to the observed
one \citep[see Figs.~3 and 5 of][]{ferrigno:11a}.
A fast-falling plasma can significantly enhance the angular difference
in beaming between the contribution from photons at the cyclotron
fundamental resonance and the overall radiation: for a static
plasma the photons at the cyclotron fundamental mostly escape
perpendicular to the field (Fig.~\ref{fig:bp0}, left) and only the
neutron star surface introduces an asymmetry with respect to the
overall radiation. For the case of bulk velocity a broader downward
beam at the cyclotron fundamental and upward beams at the cyclotron
higher harmonics are generated (Fig.~\ref{fig:bp0}, right). Hence, the
phase dependence of peaks at the fundamental and the harmonic
cyclotron features are expected to differ for an appropriate geometry,
as has been seen in 4U0115+63 \citep{tsygankov:07a}. This might also
explain the observed different energy localisation of the maximum
negative phase displacement of the pulse profile peak with respect to
the fundamental and first harmonic scattering feature in the study by
Ferrigno et al. \citep[Fig.~7 and~8 of][]{ferrigno:11a}.
\begin{figure*}\sidecaption
\centering
\includegraphics[width=12.5cm]{figures/figure4}
\caption{Color plots: flux maps for the geometry of
Fig.~\ref{fig:bp}. The additional white panels show the
corresponding energy-dependent phase shifts as calculated with
respect to the full energy pulse profile using cross-correlation
as described in detail and applied to observational data by
\citet{ferrigno:11a}. \textbf{Left:} Static plasma.
\textbf{Right:} Downward bulk velocity of $v=0.4c$. The values of
$B$, $kT_\mathrm{e}$, and $\tau$ as well as the flux scaling are
the same as in Fig.~\ref{fig:maps_obs} (see
Sect.~\ref{sect:sym}).}
\label{fig:wave}
\end{figure*}
\section{Discussion and conclusions}
The fact that we still lack basic knowledge about the coupling of the
accretion flow to the magnetosphere, which determines the location and
geometry of the emitting regions, unfortunately adds a significant
amount of freedom to theoretical predictions of pulse profiles of
X-ray pulsars. It is therefore important to identify and understand
characteristic trends in the simulations that are comparable to
observable ones to narrow down the physically reasonable
parameter space and to prepare the grounds for future proper fits of a
model to observations. Focusing on the investigation of the
formation of strong changes in the pulse profile at the cyclotron
energies, we have shown how phase shifts relative to the energy-averaged pulse profile arise naturally from the intrinsic beaming by
cyclotron resonant scattering processes for suitable geometrical
setups. The assumption of a static versus a non-static plasma has
proven to be highly relevant for the overall pulse profile morphology.
A next step will be to investigate more complex geometries, for example height-dependent velocity gradients,
temperature profiles and $B$-field geometries. Especially for the non-static case the relevance of a possible reflection component from the surface due to strong downward beaming of radiation needs to be investigated. The understanding of
scattering related, apparent phase lags allows us to separate their
physical origin from other geometric effects that can cause changes in
the morphology of the energy-dependent pulse profiles such as a
height-dependent continuum or occultations by the accretion flow.
\begin{acknowledgements}
We thank the International Space Science Institute ISSI in Bern (CH)
for granting two International Team meetings on ``The physics of the
accretion column of X-ray pulsars'', which have much inspired this
collaborative work. We also thank the Bundesministerium f\"ur
Wirtschaft und Technologie for funding through Deutsches Zentrum
f\"ur Luft- und Raumfahrt grant 50\,OR\,1113. MTW is supported by
the US Office of Naval Research and the NASA ADAP Program under
grant NNH13AV18I. We thank the anonymous referee for very
useful comments.
\end{acknowledgements}
|
\section{Introduction}
Whenever the modeling of random processes in biology, finance or physics requires to incorporate jumps, L\'evy processes are one of the building blocks under consideration. Consequently, their statistical analysis attracted much attention in the last decades. The estimation of the jump distribution, characterized by the L\'evy measure, is of particular interest. So far, only the jump density or linear functionals of it like the corresponding distribution function has been the aim estimation procedures. In the present work, we study the estimation of the (generalized) quantiles of the L\'evy measure that is the inverse of the distribution function. For a given intensity $\tau>0$ the quantile $q_\tau^+$ is the minimal jump height such that jumps larger than $q_\tau^+$ are expected only $\tau$ times in one time unit. Equivalently a jump larger than $q_\tau^+$ is expected only once in $1/\tau$ time units. $q_\tau^-$ is analogously defined for negative jumps, see Section~\ref{sec:genQuan} for the precise definitions and a discussion of possible applications.
We will consider two different observation schemes: If we directly observe the L\'evy process at equidistant discrete time points $\Delta,2\Delta,\dots ,n\Delta$, the estimator relies on the increments $L_{\Delta k}-L_{\Delta(k-1)}$, for $k=1,\dots,n$, which are independent and identically distributed according to the law of $L_\Delta$. We will focus on low-frequency observations where $\Delta>0$ remains fixed while $n\to\infty$.
The second observation scheme is motivated by an application in finance. Modeling an asset as exponential of a L\'evy process $L$, we use prices of put and call options to estimate the characteristics of the L\'evy process under the risk-neutral measure. This allows to estimate how shocks, in the sense of large jumps, are priced into the asset. Since the observed option prices are noisy, we face a statistical problem of regression type. The error analysis of the nonparametric estimators is similar to the case of low frequent direct observations.
In a high-frequency regime, i.e. observing $L_{k\Delta}$ with $\Delta\downarrow 0$, we almost see the jumps in the path and thus the estimation of the L\'evy measure is relatively straight forward, see \cite{aitSahaliaJacod:2012} for a review. As noticed by \cite{neumannReiss2009} in the low-frequency regime, i.e. $\Delta>0$ is fixed, the nonparametric estimation problem is more difficult because the number of jumps that occurred within an increment is not identifiable. Using however the L\'evy--Khintchine formula which links the characteristic function of the marginal distribution and the characteristic triplet, we can estimate the jump measure by a spectral approach. This idea was initiated by \cite{belomestnyReiss2006} and then studied further, see \cite{reiss2013} for an overview. The observation scheme of the option prices was first studied by \cite{contTankov:2004} as well as \cite{belomestnyReiss2006}.
With the notable exception of \cite{belomestny2010}, who estimates the fractional order of a L\'evy process, only linear functionals of the jump measure are considered. Instead, we solve the substantially more demanding problem of estimating the nonlinear generalized quantiles in this nonlinear inverse problem. The conditions we impose on the models are very weak. We allow the whole spectrum of L\'evy processes, reaching from diffusions to pure jump processes with unbounded variation and the combination of both. The quantile estimators are very robust in the sense that they do not depend on the drift and volatility parameters, which may be surprising. In both above described observation schemes we derive convergence rates for the quantile estimators (Theorems~\ref{thm:rateQuantiles} and \ref{thm:quantileFin}). In view of the literature our rates appear to be minimax optimal. As side results convergence rates for density estimation and distribution function estimation are obtained which are novel due to the
studied generality of L\'evy processes.
The question of adaptive estimation methods for L\'evy processes in the high-frequency and low-frequency regime has been only recently addressed by \cite{comteGenonCatalot2011} and \cite{kappus2012}, respectively, who apply model selection procedures. In the exponential L\'evy model we provide an adaptive method of Lepski-type that achieves the optimal rates (Theorem~\ref{thm:quantileFinAdapt}) with an additional $(\log \log n)$-payment for adaptivity that appears to be unavoidable. Since the exponential L\'evy model is related to the low-frequency regime let us compare our results to one by \cite{kappus2012} who has estimated linear functionals of the L\'evy density. Profiting from the regression structure, we can handle a broad class of processes while results in \cite{kappus2012} are restricted to L\'evy processes of bounded variation. The model selection approach leads to a finite sample oracle inequality where the constants, however, are not explicitly determined which might be problematic for
applications. Although our error analysis is completely asymptotic, our data-driven method achieves very good results in finite sample situations, as illustrated with prices of options on the German DAX-index.
In the next Section the generalized quantiles are introduced and their applications are discussed. In addition the basic estimation idea is outlined. Discrete observations of the L\'evy process and the option price model are considered in Sections~\ref{sec:quantileEst} and \ref{sec:FinancialQuantile}, respectively. In the latter model we construct an adaptive version of the estimator in Section~\ref{sec:lepski}. Simulations and a real data study based on DAX-options are given in Section~\ref{sec:sim}. All proofs are postponed to Section~\ref{sec:LevyProofs}.
\section{Generalized quantiles and estimation principle}\label{sec:genQuan}
Due to the L\'evy--It\^o decomposition, any L\'evy process $L=\{L_t:t\ge0\}$ can be represented as the sum of a deterministic drift determined by a parameter $\gamma\in\R$, a Brownian motion with volatility $\sigma^2>0$ and an independent jump component. The distribution of the jump sizes is described by the L\'evy measure $\nu$ which may have a singularity at zero, but satisfies $\int (x^2\wedge 1)\nu(\d x)<\infty$. The process $L$ is uniquely determined by this so-called characteristic triplet $(\sigma^2,\gamma,\nu)$.
By definition $\nu(A)$, for any Borel set $A\in\mathscr{B}(\R)$, is the expected number of jumps per unit time whose size belongs
to $A$. Taking into account the possible singularity of $\nu$ at
zero, the generalized distribution function is defined by
\begin{equation}\label{eq:distfct}
N(t):=\begin{cases}
\nu((-\infty,t]), & \text{for }t<0,\\
\nu([t,\infty)), & \text{for }t>0.
\end{cases}
\end{equation}
Our aim is to estimate its inverse function which we call \emph{generalized
quantile function}. For a given level $\tau\in(0,\nu(\R_\pm))$ we introduce
\[
q_\tau^-:=\sup\big\{t\ge0:\nu\big((-\infty,-t]\big)\ge\tau\big\}\quad\text{and}\quad
q_\tau^+:=\sup\big\{t\ge0:\nu\big([t,\infty)\big)\ge\tau\big\}.
\]
Hence, $q_{\tau}^{+}$ (resp. $q_{\tau}^{-}$) is the largest value such that jumps larger than $q_\tau^+$ (resp. smaller than $-q_\tau^-$) have a least intensity $\tau$.
Especially the distribution of large jumps, corresponding to small values of $\tau$, are conveniently described using generalized quantiles. Let us discuss a few possible applications. For L\'evy processes with compound Poisson jump component, the quantiles of the L\'evy measure, which then are a finite measures, have very similar properties as quantiles of probability distributions. The total jump intensity of negative and positive jumps is given by $q_0^-$ and $q_0^+$, respectively, and the quantiles can be used as measures of location, scale and skewness of the jump distribution.
These properties extend to infinite measures, too. For instance, the Bowley skewness for probability measures directly generalizes to $|q_\tau^--q_\tau^+|/(q_\tau^-+q_\tau^-)$ for $\tau>0$ quantifying symmetry between positive and negative jumps of a L\'evy processes. Another remarkable property is illustrated in the following example:
\begin{example}
A common construction of L\'evy measures is the so-called exponential tilting where a L\'evy measure $\nu$ is multiplied by an exponential factor, i.e., $\tilde\nu_\lambda(\d x):=e^{-\lambda |x|}\nu(\d x)$ for $\lambda>0$. Supposing $\nu$ has a Lebesgue density which is bounded outside of a neighborhood of the origin and which converges polynomially fast to zero as $|x|\to\infty$, it is easy to see that the quantiles $q^\pm_{\tau,\lambda}$ corresponding to $\tilde\nu_\lambda$ grow like $|\log \tau|/\lambda$ as $\tau\to0$. Ploting $q^\pm_{\tau,\lambda}$ against $q^\pm_{\tau,0}$, the parameter $\lambda$ thus approximately shows up as inverse of the slope for large values of $q_{\tau,0}$.
\end{example}
Therefore, estimators for the generalized quantiles are important descriptive statistics for the jump behavior of L\'evy processes. But there are further applications: To construct and estimate L\'evy copulas, the generalized quantiles are necessary, cf. \cite[Chap. 5]{contTankov:2004b} and \cite{bucherVetter2013}, and in the context of modeling prices processes, they can be used to estimate the risk within the model. One of the most popular risk measures is the value-at-risk at some level $\tau\in(0,1)$ which is given by the $(1-\tau)$-quantile of the distribution of the loss of the asset. The generalized quantiles of $\nu$ are a closely related concept which takes only the influence of shocks into account. Following this idea, the quantiles may also be useful for dynamic quantile hedging in the spirit of \cite{follmerLeukert1999}.
Now, how to estimate these quantiles? Before we rigorously introduce the estimators and study their asymptotic properties in the following two sections let us outline the general estimation principle. We follow a similar strategy as \citet{dattnerEtAl2013} who study quantile estimation in the classical deconvolution model. Compared to deconvolution, the L\'evy model is harder for two reasons. First, it is a nonlinear inverse problem such that we have to linearize the estimation error and the remainder needs extra care. Second, the underlying deconvolution problem is determined by the distribution of the process itself. Consequently, there is a strong interplay between the jump measure that we want to estimate and the underlying deconvolution operator.
Assuming $\int x^{2}\nu(\d x)<\infty$, $\phi_t$ is given by the L\'evy--Khintchine representation in Kolmogorov's version:
\begin{equation}\label{eqLevyKhint}
\phi_{t}(u):=\E[e^{iuY_{t}}]=e^{t\psi(u)}\quad\text{with}\quad\psi(u):=-\frac{\sigma^{2}}{2}u^{2}+i\gamma u+\int\big(e^{iux}-1-iux\big)\nu(\d x).
\end{equation}
Differentiating twice the characteristic exponent $\psi$, we obtain the estimating equation
\begin{equation}\label{eq:psiPP}
\psi''(u)=-\sigma^{2}-\F[x^{2}\nu](u)=\frac{\phi_{t}''(u)\phi_{t}(u)-\phi'_t(u)^2}{t\phi_t^{2}(u)}.
\end{equation}
As starting point we need an estimator of the characteristic function $\phi_t$ of the marginal distribution $L_t$ of the L\'evy process for some $t>0$. Using discrete observations of the L\'evy process, $\phi_t$ can be estimated by the empirical measure of the increments. In the financial model the characteristic function can be estimated via the pricing formula that links $\phi_t$ and the option prices. To estimate $q_\tau^\pm$, we will then apply the following program:
\begin{enumerate}
\item A density estimator for the jump measure $\nu$ can be constructed by replacing $\phi_t$ in \eqref{eq:psiPP} with its estimator, regularizing with a band limited kernel and applying the inverse Fourier transform to the left and right-hand side of \eqref{eq:psiPP}.
\item A plug-in approach yields an estimator for the distribution function.
\item The generalized $\tau$-quantiles can be estimated by minimizing the distance between the value of distribution function estimator and $\tau$.
\end{enumerate}
The estimator for the jump density is similar to the estimators proposed in \citet{nicklReiss2012} as well as \cite{kappus2012}. However, both articles are restricted to L\'evy processes with bounded variation and thus use only the first derivative of $\psi$. Focusing on settings that allow for parametric rates, the distribution function estimation have been considered by \cite{nicklReiss2012} and \cite{nicklEtAl2013} in the low and high-frequency regime, respectively.
\section{Discrete observations of the process}\label{sec:quantileEst}
We observe $n\in\mathbb{N}$ increments of the L\'evy process $L$
at equidistant time points with observation distance $\Delta>0$:
\begin{align*}
Y_{k}: & =L_{\Delta k}-L_{\Delta(k-1)},\quad k=1,\dots,n.
\end{align*}
The law of $Y_{k}$ will be denoted by $P_{\Delta}$. Using the empirical characteristic function $\phi_{\Delta,n}(u)=\frac{1}{n}\sum_{k=1}^{n}e^{iuY_{k}}$, we obtain an empirical version of $\psi''$ from \eqref{eq:psiPP}:
\begin{equation*}
\hat{\psi}_{n}''(u)=\frac{\phi_{\Delta,n}''(u)\phi_{t}(u)-\phi'_{\Delta,n}(u)^2}{t\phi_{\Delta,n}^{2}(u)}\1_{\{|\phi_{\Delta,n}(u)|\ge(\Delta n)^{-1/2}\}},
\end{equation*}
where we multiply with the indicator function to stabilize against large stochastic errors. We define the density estimator as
\[
\hat{\nu}_{h}(t):=-t^{-2}\F^{-1}\Big[\hat{\psi}_{n}''(u)\F K(hu)\Big](t),\qquad t\neq0,
\]
where $K$ is a band-limited kernel with bandwidth $h>0$ satisfying for some order $p\in\mathbb{N}$
\begin{gather}
\begin{split}
\int_{\R}K(x)\d x=1,\qquad\int x^{l}K(x)\d x=0\quad\text{ for }l=1,\dots,p,\\
\supp\F K\subset[-1,1],\quad x^{p+1}K(x)\in L^{1}(\R).\qquad
\end{split}
\label{eq:propKernel}
\end{gather}
Note that $\hat\nu_h$ depends neither on the unknown volatility $\sigma^{2}$ nor on the drift parameter $\gamma$. The distribution function can be estimated via the left and the right tail integrals
\begin{align}
\hat{N}_{h}(t) & =-\int g_{t}(x)\F^{-1}\Big[\hat{\psi}_{n}''(u)\F K(hu)\Big](x)\d x\quad\text{with\quad}g_{t}(x):=\begin{cases}
x^{-2}\1_{(-\infty,t]}, & t<0,\\
x^{-2}\1_{[t,\infty)}, & t>0.
\end{cases}\label{eq:DefDistEst}
\end{align}
Owing to $g_t\in L^1(\R)$, the estimator $\hat N_h$ is always well defined.
Assuming absolute continuity of $\nu$ at $\pm q_{\tau}^{\pm}$
with respect to the Lebesgue measure, the generalized quantiles $q_{\tau}^{+}>0$ and $q_{\tau}^{-}>0$ for a given level $\tau\in(0,\nu(\R_\pm))$ are determined by
\[
N(-q_{\tau}^{-})=\tau=N(q_{\tau}^{+}).
\]
If $\nu$ has finite mass on the negative or the positive halfline,
the $\tau$-quantiles only exist if $\tau\le\nu(\R_-)$ or $\tau\le\nu(\R_+)$, respectively. On the other hand for processes with high jump activity, i.e., $x^2\nu(x)\to\infty$ as $|x|\to0$, the bias of the distribution function estimator $\hat N_h(t)$ explodes as $|t|\to0$, cf. Proposition~\ref{prop:bias}. Since $\nu$ is not known, it is thus reasonable to estimate $q_{\tau}^{\pm}\vee\eta_{n}$
for some threshold $\eta_{n}>0$ and any $\tau>0$ instead of the quantiles themselves in order to stabilize the estimation problem.
For finite jump activity processes with regular jump densities at zero, the threshold value $\eta_{n}$ may converge slowly
to zero as $n\to\infty$ and we conclude that $\tau>\nu(\R^\pm)$ holds almost surely if $\eta_n\to0$ and the estimators take the value $\eta_{n}$ for all $n$.
Using $\hat N_h$, a minimum contrast estimation approach yields the quantile estimator
\[
\hat{q}_{\tau,h}^{\pm}:=\argmin_{t\in[\eta_{n},\infty)}|\hat{N}_{h}(\pm t)-\tau|
\]
for threshold values which are either fixed or which logarithmically decay to zero. Note that we only consider the distribution function outside a neighborhood of the origin and thus our estimator does not depend on the diffusion component which corresponds to a Dirac measure at zero with mass $\sigma^2$.
Since $\hat q_{\tau,h}$ is a point estimator, for any fixed $\tau$, nonparametric convergence rates depend on the local smoothness of $\nu$ and it is natural to assume H\"older regularity. Before specifying the exact nonparametric classes of L\'evy triplets that we will consider, let us introduce some notation. For an open set $U\subset\R$ the space of all functions continuous on $U$ is denoted by $C(U)$. The set of all functions which are H\"older-regular with index $s>0$ on $U$ is given by
\begin{align*}
C^s(U)&:=\{f\in C(U):\|f\|_{C^s(U)}<\infty\}\quad\text{with}\quad\\
\|f\|_{C^s(U)}&:=\sum_{k=0}^{\lfloor s\rfloor}\sup_{x\in U}|f^{(k)}(x)|+\sup_{x,y\in U:x\neq y}\frac{|f^{\lfloor s\rfloor}(x)-f^{(\lfloor s\rfloor)}(y)|}{|x-y|^{s-\lfloor s\rfloor}},
\end{align*}
where $\lfloor s\rfloor$ denotes the smallest integer strictly smaller than $s$. If not specified differently, $\|\cdot\|_{L^p}$ with $p\ge1$ denotes the $L^p$-norm on the whole real line with the usual notation $\|\cdot\|_\infty$ for the supremum norm. The bounded variation norm of a function $f$ on an interval $[a,b]$ for $a<b$ is defined as
\[
\|f\|_{BV([a,b])}:=\sup\Big\{\sum_{i=1}^n|f(x_i)-f(x_{i-1})|:n\in\N,a\le x_1<\dots<x_n\le b\Big\}.
\]
Define for the open set $U$, regularity $s>0$, number of moments $m>0$ and radius $R>0$ the class of L\'evy triplets
\begin{align*}
\mathcal{C}^{s}(m,U,R):= & \Big\{(\sigma^{2},\gamma,\nu)\Big|\sigma^{2}\in[0,R],\gamma\in\R,\|x^{m}\nu\|_{L^{1}}\le R,\notag\\
& \qquad\qquad\quad\nu\text{ has a Lebesgue density on }U\text{ with }\|\nu\|_{C^{s}(U)}\le R\Big\}.
\end{align*}
For the error analysis we distinguish between infinitely divisible distributions with polynomially decaying characteristic functions and exponentially decaying characteristic functions. For $\alpha,\delta,r>0$ and $\beta\in(0,2]$ set
\begin{align}
\mathcal{D}^{s}(\alpha,m,U,R):= & \Big\{(0,\gamma,\nu)\in\mathcal{C}^{s}(m,U,R)\Big|\|(1+|\bull|)^{-\Delta\alpha}/\phi_{\Delta}\|_{\infty}\le R,\|x\nu\|_{\infty}\le R,\notag\\
& \qquad\qquad\qquad\qquad\qquad\quad \|x\nu\|_{BV([-\delta,\delta])}\le R\Big\},\notag\allowdisplaybreaks\\
\mathcal{E}^{s}(\beta,m,U,r,R):= & \Big\{(\sigma^{2},\gamma,\nu)\in\mathcal{C}^{s}(m,U,R)\Big|\|\exp(-r\Delta|\bull|^{\beta})/\phi_{\Delta}\|_{\infty}\le R\Big\}.\label{eq:ClassDist}
\end{align}
We will see that the class $\mathcal{D}^{s}(\alpha,m,U,R)$ corresponds to mildly ill-posed estimation problems. As shown in \cite{trabs2014} a polynomial decay of $\phi_\Delta$ is, up to a mild regularity assumption, equivalent to $x\nu$ beeing of bounded variation near the origin and, of course, $\sigma=0$. Hence, the conditions imposed in $\mathcal D^s$ are quite natural. Moreover, they allow to apply the Fourier multiplier theorem in \cite{trabs2014}. The estimation problem in the class $\mathcal{E}^{s}(\beta,m,U,R)$ is severely ill-posed leading to logarithmic convergence rates. Because the rates are so slow, we only need very mild assumptions in $\mathcal{E}^{s}(\beta,m,U,r,R)$ without any additional condition on the jump measure. This class especially contains all infinite variation processes, noting that L\'evy processes with a diffusion component satisfy the decay condition on $\phi_\Delta$ only for $\beta=2$.
The convergence rates will be established in the $\mathcal O_P$-sense which corresponds to the loss function of confidence intervals. Since the results hold uniformly over the given classes of L\'evy processes, we define the \emph{uniform stochastic Landau symbol} $\mathcal O_{P,\Theta}$ over a parameter set $\Theta$: For random variables $(A_n)_{n\in\N}$ we write $A_n=\mathcal O_{P,\Theta}(1)$ if
\begin{equation}\label{eq:UniformOp}
\lim_{R\to\infty}\limsup_{n\to\infty}\sup_{\theta\in\Theta}P_\theta(A_n>R)=0.
\end{equation}
Since our estimation procedure relies on a plug-in approach using the density estimator $\hat \nu_h$, we start with its asymptotic behavior beeing of independent interest. As in the analysis of the quantile estimator in the deconvolution model by \cite{dattnerEtAl2013}, the following proposition is the first building block for showing the rates of $\hat q_{\tau,h}$. We will need the result for the uniform loss.
\begin{prop}\label{prop:DensLevyRate}
Let $\alpha,\beta,s,r,R>0, m>4$ and let the kernel satisfy (\ref{eq:propKernel})
with order $p\ge s$ and let $U\subset\R$ be a bounded, open set
which is bounded away from zero. Then we have for $n\to\infty$
\begin{enumerate}
\item uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$ for $h=h_{n,\Delta}=(\frac{\log n\Delta}{n\Delta})^{1/(2s+2\Delta\alpha+1)}$
\[
\sup_{t\in U}|\hat{\nu}_{h}(t)-\nu(t)|=\mathcal{O}_{P,\mathcal{D}^{s}}\Big(\Big(\frac{\log n\Delta}{n\Delta}\Big)^{s/(2s+2\Delta\alpha+1)}\Big),
\]
\item uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)$ for $h=h_{n,\Delta}=(\frac{\delta\log n\Delta}{2r\Delta})^{-1/\beta}$ with $\delta\in(0,3/2)$
\[
\sup_{t\in U}|\hat{\nu}_{h}(t)-\nu(t)
=\mathcal{O}_{P,\mathcal{E}^{s}}\Big(\Big(\frac{\log n\Delta}{\Delta}\Big)^{-s/\beta}\Big).
\]
\end{enumerate}
\end{prop}
In the mildly ill-posed case the rates correspond to the deconvolution problem with an error distribution whose characteristic function decays with polynomial rate $\Delta\alpha$. It can easily be verified that the for the pointwise loss we have the same rates without the logarithm in (i). They coincide with the convergence rates for the pointwise loss by \citet[Thm. 3.5]{kappus2012}, who has considered only finite variation L\'evy processes. \cite{kappus2012diss} have shown that these rates are minimax optimal. In the classical density estimation the logarithm is known to be unavoidable for the uniform loss. In comparison to the deconvolution model, the logarithmic rates in (ii) appear to be sharp, too.
The distribution function estimator $\hat N_h$ was studied by \citet{nicklEtAl2013} in a high-frequency regime. For low frequency observations a modification of $\hat{N}_{h}$ was considered by \citet{nicklReiss2012}. In both articles a uniform central limit theorem has been established. To this end, assumptions have been imposed which ensure that the parametric rate can be attained. Therefore, it is of interest to derive convergence rates for $\hat{N}_{h}$ in more general situations. We remark that the following result could be strengthened to the uniform loss on $\R\setminus[-\eta,\eta]$ for any $\eta>0$, cf. \eqref{eq:ConsistencyBoundLevy} below.
\begin{prop}\label{prop:DistFunct}
Let $U\subset\R$ be an open set, $t\in U$
and $\alpha,\beta,s,r,R>0$ and $m>4$. Suppose the kernel satisfies
(\ref{eq:propKernel}) with order $p\ge s+1$. Let $\Delta\in(0,1)$
and $n\to\infty$. Then:
\begin{enumerate}
\item We obtain uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$ with the bandwidth $h=h_{n,\Delta}=(n\Delta)^{-1/(2s+(2\Delta\alpha\vee1)+1)}$
\begin{align*}
|\hat{N}_{h}(t)-N(t)| & =\mathcal{O}_{P,\mathcal{D}^{s}}(r_{n,\Delta}),\quad r_{n,\Delta}:=\begin{cases}
(n\Delta)^{-(s+1)/(2s+2\Delta\alpha+1)}, & \text{for }\Delta\alpha>1/2,\\
(n\Delta)^{-1/2}(\log n\Delta)^{1/2}, & \text{for }\Delta\alpha=1/2,\\
(n\Delta)^{-1/2}, & \text{for }\Delta\alpha\in(0,1/2).
\end{cases}
\end{align*}
\item The choice $h=h_{n,\Delta}=(\frac{\delta}{2r})^{-1/\beta}(\frac{\log(n\Delta)}{\Delta})^{-1/\beta}$,
for any $\delta\in(0,3/2)$, yields uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)$
\[
|\hat{N}_{h}(t)-N(t)|
=\mathcal{O}_{P,\mathcal{E}^{s}}\Big(\Big(\frac{\log(n\Delta)}{\Delta}\Big)^{-(s+1)/\beta}\Big).
\]
\end{enumerate}
\end{prop}
As expected, the convergence rates for $\hat N_n$ are faster than for density estimation because we gain one degree of smoothness. In particular, we achieve the parametric rate for a L\'evy process with very slowly decaying characteristic function or for $\Delta$ sufficiently small.
Following the standard M-estimation strategy, we use
a Taylor expansion to analyze the estimation error of the quantile estimators. Using that $\hat{N}_{h}'(t)=-\sign(t)\hat{\nu}_{h}(t)$ for $t\neq0$, we obtain
\begin{align*}
0\approx\hat{N}_{h}(\hat{q}_{\tau,h}^{+})-\tau=\hat{N}_{h}(q_{\tau}^{+})-N(q_{\tau}^{+})-(\hat{q}_{\tau,h}^{+}-q_{\tau}^{+})\hat{\nu}_{h}(\xi^+)
\end{align*}
for some intermediate point $\xi^+$ between $q_{\tau}^{+}$and $\hat{q}_{\tau,h}^{+}$
and similarly for $\hat{q}_{\tau,h}^{-}$. By continuity of $\hat N_h$ and the construction of $\hat q_{\tau,h}^\pm$ the probability of the event $\{\hat{N}_{h}(\hat{q}_{\tau,h}^{+})-\tau=0\}$ converges to one, cf. \eqref{eq:ZEst}. On this event the estimation error can therefore be represented as
\begin{equation}
|\hat{q}_{\tau,h}^{\pm}-q_{\tau} ^{\pm}|=\Big|\frac{\hat{N}_{h}(\pm q_{\tau}^{\pm})-N(\pm q_{\tau}^{\pm})+\tau-\hat{N}_{h}(\hat{q}_{\tau,h}^{+})}{\hat{\nu}_{h}(\xi^{\pm})}\Big|=\frac{|\hat{N}_{h}(\pm q_{\tau}^{\pm})-N(\pm q_{\tau}^{\pm})|}{|\hat{\nu}_{h}(\xi^{\pm})|},\label{eq:errorRep}
\end{equation}
for intermediate points $\xi^\pm$. If the denominator does not explode, $q_{\tau}^{\pm}$ can be estimated with the same rate as $N(\pm q_\tau^\pm)$. Since the convergence rates for the density estimator $\hat\nu_h$ and for the distribution function estimator $\hat{N}_{h}$ are already established, it remains to show consistency of the quantile estimator $\hat{q}_{\tau,h}^{\pm}$ itself. To this end, a minimal global regularity of $\nu$ is required. Note that assuming a bounded density would be far to restrictive for L\'evy measures.
We need to specify our nonparametric classes further such that the
quantiles exist. Writing $\nu\in C^{s'}(\R\setminus[-\eta,\eta])$ for $s'\in(-1,0],\eta>0$ if the antiderivatives $y\mapsto\int_{-\infty}^{-y} \nu(\d x)$ and $y\mapsto\int^\infty_{y} \nu(\d x)$ are in $C^{1+s'}((\eta,\infty))$, we define for a given $\tau>0$ and with $\zeta>0$
\begin{align}
&\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,m,\zeta,\eta,R)\label{eq:Dtilde}\\
&\quad:= \Big\{(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}\big(\alpha,m,(q_{\tau}^{+}-\zeta,q_{\tau}^{+}+\zeta)\cup(q_{\tau}^{-}-\zeta,q_{\tau}^{-}+\zeta),R\big)\Big|\notag\\
&\qquad\qquad\exists q_{\tau}^{+},q_{\tau}^{-}\in(\eta,\infty): N(-q_{\tau}^{-}) =\tau=N(q_{\tau}^{+}), \|\nu\|_{C^{s'}(\R\setminus[-\eta,\eta])}<R, \nu(q_{\tau}^{\pm})>\tfrac{1}{R}\Big\},\notag\\
&\mathcal{\tilde{E}}_{\tau}^{s,s'}(\beta,m,\zeta,\eta,r,R)\label{eq:ClassQuantile}\\
&\quad:= \Big\{(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}\big(\beta,m,(q_{\tau}^{+}-\zeta,q_{\tau}^{+}+\zeta)\cup(q_{\tau}^{-}-\zeta,q_{\tau}^{-}+\zeta),r,R\big) \Big|\notag\\
&\qquad\qquad\exists q_{\tau}^{+},q_{\tau}^{-}\in(\eta,\infty):N(-q_{\tau}^{-})=\tau=N(q_{\tau}^{+}), \|\nu\|_{C^{s'}(\R\setminus[-\eta,\eta])}<R, \nu(q_{\tau}^{\pm})>\tfrac{1}{R}\Big\}.\notag
\end{align}
The condition $\nu(q_{\tau}^{\pm})>\tfrac{1}{R}$ especially implies that the quantiles are unique. As expected from the representation \eqref{eq:errorRep}, we obtain the same rates for quantile estimation as for distribution function estimation.
\begin{theorem}\label{thm:rateQuantiles}
Let $\tau>0$ and $\alpha,\beta,s,\zeta,r,R>0,s'\in(-1,0]$
and $m>4$. Suppose the kernel satisfies (\ref{eq:propKernel}) with
order $p\ge s+1$ and $\eta_{n}>0$ with $\eta_{n}^{-1}\lesssim\log n$.
Then we obtain for $\Delta\in(0,1)$ and $n\to\infty$:
\begin{enumerate}
\item The bandwidth $h=h_{n,\Delta}=(n\Delta)^{-1/(2s+(2\Delta\alpha\vee1)+1)}$ yields uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,m,\zeta,\eta_{n},R)$
\begin{align*}
|\hat{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}| & =\mathcal{O}_{P,\mathcal{\tilde{D}}_{\tau}^{s,s'}}(r_{n,\Delta}),\quad r_{n,\Delta}:=\begin{cases}
(n\Delta)^{-(s+1)/(2s+2\Delta\alpha+1)} & \text{for }\Delta\alpha>1/2,\\
(n\Delta)^{-1/2}(\log n\Delta)^{1/2} & \text{for }\Delta\alpha=1/2,\\
(n\Delta)^{-1/2} & \text{for }\Delta\alpha\in(0,1/2).
\end{cases}
\end{align*}
\item The choice $h=h_{n,\Delta}=(\frac{\delta}{2r})^{-1/\beta}(\frac{\log(n\Delta)}{\Delta})^{-1/\beta}$,
for any $\delta\in(0,2/3)$, yields uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{\tilde{E}}_{\tau}^{s,s'}(\beta,m,\zeta,\eta_{n},r,R)$
\[
|\hat{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}|=\mathcal{O}_{P,\mathcal{\tilde{E}}_{\tau}^{s,s'}}\Big(\Big(\frac{\log(n\Delta)}{\Delta}\Big)^{-(s+1)/\beta}\Big)
\]
\end{enumerate}
\end{theorem}
\begin{remark}
In the parametric regime., i.e., $(0,\gamma,\nu)\in\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,m,\zeta,\eta_{n},R)$ with $\Delta\alpha\in(0,1/2)$, we could hope for more. In view of the central limit theorem by \cite{nicklReiss2012} a more precise analysis of the main stochastic error term, defined in \eqref{eq:DecompLin}, may lead to a central limit theorem of the distribution function estimator $\hat N_{h}$. Since the our analysis of \eqref{eq:errorRep} reveals that
\[
\hat{q}_{\tau,h}^{+}-q_{\tau}^+=\frac{\hat{N}_{h}(q_{\tau}^{+})-N(q_{\tau}^{+})+o_P(n^{-1/2})}{\nu(q_\tau^+)+o_P(1)},
\]
a central limit theorem for $\hat q_\tau^\pm$ immediately follows. Note that optimality of the asymptotic variance by \cite{nicklReiss2012} in the sense of semi-parametric efficiency is proved in \cite{trabs2013}. Using the $\Delta$-method, this lower bound can be extend to a lower bound for the asymptotic variance for the quantile estimation.
\end{remark}
\begin{remark}
For high-frequency data, that is for $\Delta\to0$ and $n\Delta\to\infty$, we obtain in $\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,\zeta,\eta_{n},R)$
always the parametric rate $(n\Delta)^{-1/2}$. In $\mathcal{\tilde{E}}_{\tau}^{s,s'}(\beta,\zeta,\eta_{n},r,R)$ the bandwidth
$h_{n,\Delta}=\Delta^{1/\beta\wedge 3/4}$ should be chosen instead such that our estimates in the proof of Theorem~\ref{thm:rateQuantiles} yield the almost optimal rate $(n\Delta)^{-1/2}|\log\Delta|$ provided that
$s$ is large enough such that the bias condition $h^{s+1}=\mathcal{O}((n\Delta)^{-1/2})$ is satisfied. With stronger assumptions the latter restriction can be circumvented and the rate can be improved. In fact, in \citet{nicklEtAl2013} we show that the parametric rate $(n\Delta)^{-1/2}$ can be obtained with this estimator under suitable conditions on $\nu$.
\end{remark}
\section{Observation of option prices}\label{sec:FinancialQuantile}
Let us consider the exponential L\'evy model for asset prices
\begin{equation}
S_{t}=S_{0}e^{rt+L_{t}},\quad t\ge0,\label{eq:expLevyMod}
\end{equation}
with initial value $S_{0}>0$, riskless interest rate $r\ge0$ and with the driving L\'evy process $L$ whose characteristic triplet is $(\sigma^{2},\gamma,\nu)$. Since L\'evy processes are a quite large and flexible class of stochastic processes, important stylized facts of financial data can be reproduced, see \cite{contTankov:2004b} for properties and examples of the exponential L\'evy model. At the same time the well understood probabilistic structure of L\'evy processes allows to construct procedures to fit the model to real data. The nonparametric calibration of this model was studied by \cite{contTankov:2004,belomestnyReiss2006} as well as \cite{trabs:2011}. \cite{soehl2014} have derived confidence sets. An empirical study of the calibration methods can be found in \cite{soehlTrabs2014}.
By arbitrage arguments the price process of an asset should be a martingale implying $\E[S_t]=S_0$ for all $t>0$. Assuming again $\int x^2\nu(\d x)<\infty$, this is equivalent to the martingale condition
\begin{equation}\label{eqMartCond}
\frac{\sigma^2}{2}+\gamma+\int_{-\infty}^\infty(e^x-1-x)\nu(\d x)=0,
\end{equation}
which we assume throughout this section.
Since we want to estimate L\'evy measure under the risk-neutral measure, the procedure is based on option prices. More precisely, we observe prices of vanilla options. Let us fix a maturity $T>0$, measured in years, and define the negative log--moneyness $x:=\log(K/S_0)-rT$ as a logarithmic transform of the strike prices $K>0$. In terms of $x$, call and put prices in model \eqref{eq:expLevyMod} are given by $\mathcal{C}(x,T)=S_0\mathbb E[(e^{L_T}-e^x)_+]$ and $\mathcal{P}(x,T)=S_0\mathbb E[(e^x-e^{L_T})_+]$, respectively. The prices can be summarized in the \emph{option function}
\begin{equation}\label{eq:optionFct}
O(x):=\begin{cases}
\displaystyle S_0^{-1}\mathcal{C}(x,T), & \quad x\ge0,\\
\displaystyle S_0^{-1}\mathcal{P}(x,T), & \quad x<0.
\end{cases}
\end{equation}
We observe $O$ at a finite number of (transformed) strike prices $x_1,\dots,x_n$, for $n\in\N$, corrupted by noise
\begin{equation}\label{eq:Observations}
O_{j}=O(x_{j})+\delta_{j}\eps_j\quad j=1,\dots,n,
\end{equation}
where $(\eps_{j})$ are i.i.d. centered random variables with $\Var(\eps_j)=1$ and with local
noise levels $(\delta_j)$. The observation errors are due to the bid--ask spread and other market frictions. By interpolating the observations $(x_{j},O_{j})_{j=1,\dots,n}$ using B-splines, we construct an empirical version $\tilde{O}$ of the option function as in \cite{belomestnyReiss2006}. From the theoretical perspective linear splines are sufficient, but in applications B-splines of higher degrees may lead to better results, cf. \cite{soehlTrabs2014} for details. Since the function $O$ is related to the characteristic function $\phi_T$ of $L_{T}$ via the pricing formula, cf. \cite{carrMadan:1999},
\begin{equation}\label{pricingFormula}
\mathcal{F}O(u):=\int_{-\infty}^{\infty}e^{iux}O(x)\d x=\frac{1-\phi_T(u-i)}{u(u-i)},
\end{equation}
we obtain an estimator of $\phi_T$ given by
\begin{equation}\label{eq:phiTilde}
\tilde\phi_{T,n}(u):=1-u(u+i)\F \tilde O(u+i).
\end{equation}
Using $\tilde{\phi}_{T,n}$, we obtain a quantile estimator as described in Section~\ref{sec:genQuan}:
\begin{enumerate}
\item The second derivative of the characteristic exponent can be estimated
by differentiating \eqref{eq:phiTilde} twice. We define
\iffalse
\begin{align*}
\psi(u) & =\frac{1}{T}\log\big(1-u(u+i)\F\mathcal{O}(u+i)\big)\\
\psi'(u) & =-\frac{(2u+i)\F\mathcal{O}(u+i)+u(iu-1)\F[x\mathcal{O}](u+i)}{T(1-u(u-i)\F\mathcal{O}(u+i))}\\
\psi''(u) & =-\frac{(2\F\mathcal{O}(u+i)+(4iu-2)\F[x\mathcal{O}](u+i)-(u^{2}+iu)\F[x^{2}\mathcal{O}](u+i))}{T(1-u(u-i)\F\mathcal{O}(u+i))}\\
& \qquad-\frac{1}{T}\Big(\frac{(2u+i)\F\mathcal{O}(u+i)+u(iu-1)\F[x\mathcal{O}](u+i)}{(1-u(u-i)\F\mathcal{O}(u+i))}\Big)^{2}
\end{align*}
\fi
\begin{align}
\tilde\psi_n'(u) & :=-\frac{(2u+i)\F\tilde{O}(u+i)+u(iu-1)\F[x\tilde{O}](u+i)}{T(1-u(u-i)\F\tilde{O}(u+i))},\notag\\
\tilde{\psi}_{n}''(u)
& :=-\frac{(2\F\tilde{O}(u+i)+(4iu-2)\F[x\tilde{O}](u+i)-(u^{2}+iu)\F[x^{2}\tilde{O}](u+i))}{T(1-u(u-i)\F\tilde{O}(u+i))}\notag\\
& \qquad-\frac{1}{T}\Big(\frac{(2u+i)\F\tilde{O}(u+i)+u(iu-1)\F[x\tilde{O}](u+i)}{1-u(u-i)\F\tilde{O}(u+i)}\Big)^{2}.\label{eq:psiTildePP}
\end{align}
Using a kernel $K$ with bandwidth $h>0$ satisfying \eqref{eq:propKernel}, we obtain the density estimator
\[
\tilde{\nu}_{h}(t):=-t^{-2}\F^{-1}\Big[\tilde{\psi}_{n}''(u)\F K(hu)\Big](t),\qquad t\neq0.
\]
\item Integrating $\tilde\nu_h$, the estimator of the generalized distribution function
is given by
\[
\tilde{N}_{h}(t)=-\int g_{t}(x)\F^{-1}\Big[\tilde{\psi}_{n}''(u)\F K(hu)\Big](x)\d x,\quad t\neq0,
\]
with $g_{t}$ from (\ref{eq:DefDistEst}).
\item For $\tau>0$ the quantile estimators are defined as the minimum contrast estimators
\[
\tilde{q}_{\tau,h}^{\pm}:=\argmin_{t\in[\eta_{n},\infty)}|\tilde{N}_{h}(\pm t)-\tau|
\]
with threshold value $\eta_{n}\downarrow0$.
\end{enumerate}
These estimators are well defined on the event $A=\{\forall u\in[-1/h,1/h]:\tilde\phi_{T,n}(u)\neq0\}$ whose probability increases if $\tilde\phi_{T,n}$ concentrates around the true $\phi_T$. In an idealized model \cite{soehl:2010} shows $P(A)=1$. Note that $\tilde \nu_h$ is different from the finite activity estimator by \citet{belomestnyReiss2006}. The k-function estimator from \citet{trabs:2011} in the self-decomposable model relies on a similar idea but using only the first derivative of $\psi$.
Since we want to concentrate on the main aspects in the error analysis and to avoid technicalities, we will work in the idealized Gaussian white noise model which was considered by \citet{soehl2014}
as well. Assume that the noise levels of the observations \eqref{eq:Observations} are given by the values $\delta_j=\delta(x_j)$, $j=1,\dots,n$, of some function $\delta:\mathbb{R}\to\mathbb{R}_{+}$. The observed strike prices are assumed to be the quantiles $x_j=F^{-1}(j/(n+1))$, $j=1,\dots,n$, of a distribution with distribution function $F:\mathbb{R}\to[0,1]$ and density $f>0$. Incorporating the observation errors as well as their distribution, we define the general noise level
\begin{equation}\label{varrho}
\varrho(x)=\delta(x)/\sqrt{f(x)}.
\end{equation}
For standard normal $(\eps_j)$ \cite{brownLow1996} have shown asymptotic equivalence in the sense of Le Cam of the nonparametric regression model~\eqref{eq:Observations} and the Gaussian white noise model
$$\d Z(x)=O(x)\d x+n^{-1/2}\varrho(x)\d W(x)$$
with a two-sided Brownian motion $W$ for $x$ on a possibly growing bounded interval. The equivalence to more general error distributions follows from \cite{gramaNussbaum:2002}. More details on this equivalence can be found in \cite{soehl2014} and \citet[Supplement]{trabs:2011}.
$Z$ is an empirical version of the antiderivative of $O$. In that sense we define $\F\tilde{O}(u):=\F[\d Z](u)=\F O(u)+n^{-1/2}\int e^{iux}\varrho(x)\d W(x)$ and analogously for $\F[x\tilde O]$ and $\F[x^2\tilde O]$. Owing to \eqref{eq:phiTilde}, the estimation
error of the $\tilde{\phi}_{T,n}$ is given by the Gaussian process
\[
\Phi_{n}(u):=(\tilde{\phi}_{T,n}-\phi_{T})(u)=u(u+i)\F[O-\tilde{O}](u+i)=n^{-1/2}u(u-i)\int e^{iux-x}\rho(x)\d W(x).
\]
While in the previous section the concentration of the estimator $\phi_{\Delta,n}$ around $\phi_\Delta$ was obtained by the i.i.d. structure of the increments, we use here the concentration of Gaussian measures. Applying Dudley's entropy theorem, we obtain the following path property of
$\Phi_n$. This lemma is in line with the results by \citet{soehl:2010}
and Proposition~1 in \citet{soehl2014}.
\begin{lemma}
\label{lem:GaussProc}Grant $\int(1+|x|)^{m}e^{-2x}\rho^{2}(x)\d x<\infty$
for some $m>4$. Then $\Phi_n$ is twice $L^{2}(P)$-differentiable
with derivatives
\begin{align}
\Phi_{n}^{(1)}(u):= & n^{-1/2}\int e^{iux-x}\big(2u-i+(iu^{2}+u)x\big)\rho(x)\d W(x),\label{eq:GaussProc}\\
\Phi_{n}^{(2)}(u):= & n^{-1/2}\int e^{iux-x}\big(2+2(2iu+1)x+u(i-u)x^{2}\big)\rho(x)\d W(x).\nonumber
\end{align}
Moreover, $\Phi_{n}^{(0)}:=\Phi_{n},\Phi_{n}^{(1)}$ and $\Phi_{n}^{(2)}$ have versions
that are almost surely continuous and satisfy for any $U>0$
\[
\E\big[\|\Phi_{n}^{(k)}\|_{L^{\infty}[-U,U]}\big]=\mathcal O\big( n^{-1/2}U^{2}\sqrt{\log U}\big)\quad\text{for}\quad k=0,1,2.
\]
\end{lemma}
In the following we may use these almost surely continuous and bounded versions.
Let us first state a result on the uniform loss for the density estimator
$\tilde{\nu}_{h}$.
\begin{prop}
\label{prop:UniformDensRates} Let $\alpha,\beta,s,r,R>0$, suppose the
kernel satisfies (\ref{eq:propKernel}) with order $p\ge s$ and let
$U\subset\R$ be a bounded, open set which is bounded away from zero.
Suppose $\|(1\vee x^{2})e^{-x}\rho(x)\|_{\infty}<\infty$. Then we
have for $n\to\infty$
\begin{enumerate}
\item with $h=h_{n}=\big(\frac{\log n}{n}\big)^{1/(2s+2T\alpha+5)}$ uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,2,U,R)$ satisfying \eqref{eqMartCond}
\begin{align*}
\sup_{t\in U}|\tilde{\nu}_{h}(t)-\nu(t)| & =\mathcal{O}_{P,\mathcal D^s}\Big(\Big(\frac{\log n}{n}\Big)^{s/(2s+2T\alpha+5)}\Big),
\end{align*}
\item with $h=h_{n}=(\frac{1}{4r}\log n)^{-1/\beta}$ uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,2,U,r,R)$ satisfying \eqref{eqMartCond}
\[
\sup_{t\in U}|\tilde{\nu}_{h}(t)-\nu(t)|=\mathcal{O}_{P,\mathcal E^s}\big((\log n)^{-s/\beta}\big).
\]
\end{enumerate}
\end{prop}
The convergence rates for the pointwise loss are the same without the logarithmic factor in (i). They coincide with the rates by \cite{belomestnyReiss2006} who have considered only the extreme cases: If $\sigma^2=0$ and $\nu$ is a finite measure, the pointwise risk converges with rate $n^{-s(2s+5)}$, and if $\sigma^2>0$, we obtain the rate $(\log n)^{-s/2}$. The estimator for the k-function by \cite{trabs:2011} achieves the same rate as the corresponding pointwise result in Proposition~\ref{prop:UniformDensRates}(i). Since in the two afore mentioned papers lower bounds have been proved and the logarithm is unavoidable for uniform loss, the above rates appear to be minimax optimal.
Recalling the function classes from \eqref{eq:Dtilde}, we obtain the following convergence rates for the quantile estimators $\tilde{q}_{\tau,h}^{\pm}.$
\begin{theorem}
\label{thm:quantileFin}Let $\tau>0$ and $\alpha,\beta,s,\zeta,r,R>0,s'\in(-1,0]$.
Suppose $\|(1\vee x^{2})e^{-x}\rho\|_{\infty}\lesssim1$ and $\int(1+|x|)^{m}e^{-2x}\rho^{2}(x)\d x<\infty$ for some $m>4$. Suppose the kernel satisfies (\ref{eq:propKernel})
with order $p\ge s+1$ and let $\eta_{n}>0$ with $\eta_{n}^{-1}\lesssim\log n$.
Then we obtain for $n\to\infty$:
\begin{enumerate}
\item with $h=h_{n}=n^{-1/(2s+2T\alpha+5)}$ uniformly in $\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,2,\zeta,\eta_n,R)$ satisfying \eqref{eqMartCond}
\begin{align*}
|\tilde{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}| & =\mathcal{O}_{P,\mathcal{\tilde{D}}_{\tau}^{s,s'}}(n^{-(s+1)/(2s+2T\alpha+5)}),
\end{align*}
\item with $h=h_{n}=(\frac{1}{4r}\log n)^{-1/\beta}$ uniformly in $\mathcal{\tilde{E}}_{\tau}^{s,s'}(\beta,2,\zeta,\eta_n,r,R)$ satisfying \eqref{eqMartCond}
\[
|\tilde{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}|=\mathcal{O}_{P, \mathcal{\tilde{E}}_{\tau}^{s,s'}}\big((\log n)^{-(s+1)/\beta}\big).
\]
\end{enumerate}
\end{theorem}
Compared to Theorem~\ref{thm:rateQuantiles}, the rates in (i) are always slower. In particular, the parametric rate can never be achieved. Heuristically, this is because we estimate a derived parameter of the state price density, which is basically the second derivative of the observed option function $O$. In the Fourier domain we see in the pricing formula~\eqref{pricingFormula} that $\F O$ decays two polynomial degrees faster than $\phi_T$ such that the ill-posedness of the statistical problem is larger. In the severely ill-posed case (ii) the rate is the same in both observation schemes since the rates are only logarithmically slow. The moment assumption is weakened to second moments, which are necessary for the identification identity~\eqref{eq:psiPP}. Instead the existence of fourth moments is implicitly imposed on the error distribution in the regression scheme. Although the theorem is stated for $\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,2,\zeta,\eta,R)$, the bounded variation of $x\nu$ is not needed here and could be dropped.
\section{Data-driven choice of the bandwidth}\label{sec:lepski}
Of course the optimal bandwidth is not known to the practitioner.
To provide an adaptive method, we apply the approach by \cite{Lepski1990} by considering
a family of quantile estimators $\{\tilde{q}_{\tau,h}:h\in\mathcal{B}_{n}\}$
for an appropriate set of bandwidths $\mathcal{B}_{n}$. Following the
construction in \cite{dattnerEtAl2013}, we define for a constant
$L>1$ and a sequence $(N_{n})\subset\mathbb{N}$ satisfying $n^{-1}L^{N_{n}}\sim(\log n)^{-5}$
\[
h_{n,j}:=n^{-1}L^{j}\quad\text{ for }\quad j=0,\dots,N_{n}.
\]
To ensure that the density estimator $\tilde{\nu}_{h}$ is consistent
for any $h\in\mathcal{B}_{n}$, we choose the minimal bandwidth via
\begin{align*}
\tilde{j}_{n}:=&\min\Big\{ j=0,\dots,N_{n}:
&\frac{1}{2}\le \frac{(\log n)^{2}}{n^{1/2}}\Big(\int_{-1/h_{n,j}}^{1/h_{n,j}}\frac{(1+u^{4})\1_{\{|\tilde{\phi}_{T,n}(u)|\ge(1+|u|)^2/n^{1/2}\}}}{|\tilde{\phi}_{T,n}(u)|^{2}}\d u\Big)^{1/2}\le1\Big\}
\end{align*}
and define
\[
\mathcal{B}_{n}:=\{h_{n,\tilde{j}_{n}},\dots,h_{n,N_{n}}\}.
\]
To choose the bandwidth from $\mathcal B_n$ which mimics the oracle bandwidth in Theorem~\ref{thm:quantileFin}, we have to estimate the standard deviation of the stochastic error. This problem is similar to the one considered by \cite{soehl2014} who has determined the asymptotic distribution of the finite activity estimators by \cite{belomestnyReiss2006} and has derived confidence sets. In fact, we will only estimate an upper bound of the standard deviation. This bound should be as sharp as possible to allow for a good finite sample behavior. The main problem is that any upper bound depends on unknown quantities which have to be estimated as well with a sufficiently fast convergence rate. Our bound will not depend on any asymptotic or function class specific constants. The stochastic error is dominated by its linearization and thus we define, cf. Lemma~\ref{lem:LinFin},
\begin{align}
\tilde\Sigma_{n,h}^\pm:=&\frac{1}{2\pi n^{1/2}T}\Big(\|x^{2}e^{-x}\rho(x)\|_{\infty}\|\tilde\chi_{\tilde q_{\tau,h}^\pm}^{(2)}\|_{L^2}
+\|xe^{-x}\rho(x)\|_{\infty}\|\tilde\chi_{\tilde q_{\tau,h}^\pm}^{(1)}\|_{L^2}+\|e^{-x}\rho(x)\|_{\infty}\|\tilde\chi_{\tilde q_{\tau,h}^\pm}^{(0)}\|_{L^2}\Big)\label{eq:SigmaTilde}
\end{align}
with auxiliary functions, for $u\in\R,t\neq0$,
\begin{align*}
\tilde\chi_t^{(0)}(u)&:=\F g_{t}(-u)\F K(hu)\Big(u(u-i)\frac{T^2\tilde\psi_n'(u)^2-T\tilde\psi_n''(u)}{\tilde\phi_{T,n}(u)}+2T(i-2u)\frac{\tilde\psi_n'(u)}{\tilde\phi_{T,n}(u)}+2\tilde\phi_{T,n}^{-1}(u)\Big),\\
\tilde\chi_t^{(1)}(u)&:=\F g_{t}(-u)\F K(hu)\Big((4iu+2)\tilde\phi_{T,n}^{-1}(u)-2Tu(iu+1)\frac{\tilde\psi_n'(u)}{\tilde\phi_{T,n}(u)}\Big),\\
\tilde\chi_t^{(2)}(u)&:=u(i-u)\F g_{t}(-u)\F K(hu)\tilde\phi_{T,n}^{-1}(u).
\end{align*}
Note that $\tilde{\Sigma}_{n,h}^\pm$ are monotone decreasing in $h$. The magnitude of the stochastic error of $\tilde{q}_{\tau,h}^{\pm}$
can then be estimated by
\begin{equation}\label{eq:V}
\tilde{V}_{n}^{\pm}(h):=\frac{(1+\delta)\sqrt{2\log\log n}\tilde{\Sigma}^\pm_{n,h}}{|\tilde{\nu}_{h}(\tilde{q}_{\tau,h}^{\pm})|}
\end{equation}
for any small $\delta>0$. Defining
\[
\mathcal{U}_{h}^{\pm}:=[\tilde{q}_{\tau,h}^{\pm}-\tilde{V}_{n}^{\pm}(h),\tilde{q}_{\tau,h}^{\pm}+\tilde{V}_{n}^{\pm}(h)],
\]
the adaptive estimator is defined as
\[
\tilde{q}_{\tau}^{\pm}:=\tilde{q}_{\tau,\tilde{h}^{\pm}}^{\pm}\quad\text{with}\quad\tilde{h}^{\pm}:=\max\{h\in\mathcal{B}_{n}:\bigcap_{\mu\le h,\mu\in\mathcal{B}_{n}}\mathcal{U}_{h}^{\pm}\neq\emptyset\}.
\]
\begin{theorem}
\label{thm:quantileFinAdapt}Let $\tau>0$ and $\alpha,\beta,s,\zeta,r,R>0,s'\in(-1,0]$.
Suppose $\|(1\vee x^{2})e^{-x}\rho\|_{\infty}\lesssim1$ and $\int(1+|x|)^{m}e^{-2x}\rho^{2}(x)\d x<\infty$ for some $m>4$. Suppose the kernel satisfies (\ref{eq:propKernel})
with order $p\ge s+1$ and let $\eta_{n}>0$ with $\eta_{n}^{-1}\lesssim\log n$.
Then we obtain for $n\to\infty$:
\begin{enumerate}
\item uniformly in $(\sigma^2,\gamma,\nu)\in\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,2,\zeta,\eta_n,R)$ satisfying \eqref{eqMartCond}
\begin{align*}
|\tilde{q}_{\tau}^{\pm}-q_{\tau}^{\pm}| & =\mathcal{O}_{P,\mathcal{\tilde{D}}_{\tau}^{s,s'}}\big(((\log\log n)n^{-1})^{(s+1)/(2s+2T\alpha+5)}\big),
\end{align*}
\item uniformly in $(\sigma^2,\gamma,\nu)\in\mathcal{\tilde{E}}_{\tau}^{s,s'}(\beta,2,\zeta,\eta_n,r,R)$ satisfying \eqref{eqMartCond}
\[
|\tilde{q}_{\tau}^{\pm}-q_{\tau}^{\pm}|=\mathcal{O}_{P,\mathcal{\tilde{E}}_{\tau}^{s,s'}}\big((\log n)^{-(s+1)/\beta}\big).
\]
\end{enumerate}
\end{theorem}
In the mildly ill-posed case (i) the adaptive method looses a $\log\log n$-factor compared to the oracle choice in Theorem~\ref{thm:quantileFin}. Note that our loss function is bounded thus we loose only a $\log\log n$-factor instead of the $\log n$-factor, which would appear, for instance, for the mean squared error. In view of \cite{spokoiny1996} this payment for adaptivity is unavoidable. In the severely ill-posed case (ii) the rates are already logarithmically slow such that the adaptive method causes no additional loss.
\section{Simulations and real data }\label{sec:sim}
We will illustrate the quantile estimation method in simulations from the CMGY model introduced by \cite{Carr:2002}. The driving L\'evy process of the asset is tempered stable and may have a diffusion component. For parameters $C>0,M,G\ge0$ and $Y<2$ the L\'evy measure in the CMGY model is given by the Lebesgue density
\[
\nu_{cgmy}(x)=\begin{cases}
C|x|^{-1-Y}e^{-G|x|},\quad &x<0,\\
Cx^{-1-Y}e^{-Mx},\quad &x>0.
\end{cases}
\]
With a fifth parameter $\sigma\ge0$ the characteristic triplet of the underlying L\'evy process is $(\sigma^2,\gamma,\nu_{cgmy})$ where the drift is determined by the martingale condition \eqref{eqMartCond}. For the simulations we set $C=1, G=5, M=8, Y=0.5$ and $\sigma=0.1$ which appears to be a realistic choice in view of the empirical results by \cite{Carr:2002}. The riskless interest rate is chosen as $r=0.06$. The design points $x_j,j=1,\dots,n,$ are constructed as $j/(n+1)$-quantiles of a $\mathcal N(0,1/2)$ distribution. We simulate $n=100$ option prices with time to maturity $T=0.25$ corresponding to three months. According to a rule of thumb by \citet[cf.][p. 439]{contTankov:2004b}, the local noise levels $\delta_j$ are chosen as 1\% of the observed prices $O(x_j)$ which is later assumed for the real data, too.
\begin{table}\centering
\begin{tabular}{ccccccc}
\hline\hline
\multicolumn{3}{c}{$10^2\cdot$RMSE}&\multicolumn{2}{c}{$\tilde q_\tau^-$} & \multicolumn{2}{c}{$\tilde q_\tau^+$}\\
$\tau$ & $q_\tau^-$ & $q_\tau^+$ & oracle & adaptive & oracle & adaptive \\\hline
0.5 & 0.1778 & 0.1241 & 0.346 & 4.806 & 0.444 & 2.246\\
1.0 & 0.1201 & 0.0868 & 0.297 & 1.396 & 0.361 & 0.741\\
1.5 & 0.0929 & 0.0665 & 0.185 & 0.890 & 0.434 & 0.869\\
2.0 & 0.0726 & 0.0563 & 0.275 & 0.867 & 0.314 & 0.670\\
2.5 & 0.0624 & 0.0461 & 0.233 & 0.652 & 0.424 & 0.694\\\hline\hline
\end{tabular}
\caption{Empirical RMSE (multiplied by 100) of the quantile estimators $\tilde q_\tau^\pm$ from 1000 Monte Carlo simulations of the CGMY model.}\label{tab:cgmy}
\end{table}
In order to apply the estimation procedure, we have to choose some parameters. The truncation value is set to $\eta=0.02$. To construct the bandwidth set $\mathcal B_n$, we take $L=1.1$. To compute $\tilde \Sigma_{n,h}^{\pm}$, we need the noise function $\varrho$ from \eqref{varrho} which depends on the unknown density $f$ of the distribution of the strikes. It can be estimated from the observation points $(x_j)_{j=1,\dots,n}$ using some standard density estimation method. As in \cite{soehlTrabs2014} we will apply a triangular kernel estimator, where the bandwidth is chosen by Silverman's rule of thumb.
To assess the performance of the estimation procedure, we compare the adaptive choice of the bandwidth to the oracle bandwidth, meaning that $h$ is chosen such that the empirical root mean squared error (RMSE) is minimized. Our simulation results are summarized in Table~\ref{tab:cgmy} for $\tau\in\{0.5,1.0,1.5,2.0,2.5\}$. Although the RMSE of the adaptive method is larger than the oracle choice, the method achieves reasonable estimation errors. Note that the sample size is relatively small. For $\tau=0.5$ the RMSE is about 2.7\% of the true quantile for $\tilde q_{0.5}^-$ and 1.8\% for $\tilde q_{0.5}^+$. For larger values of $\tau$ the RMSE decays to an order of 1\% of the true quantiles. The reason is that small values of $\tau$ correspond to rare large jumps such that the jump density is small. Consequently, the estimation error \eqref{eq:errorRep} is large. Since the stochastic estimation error has to be estimated by $\tilde V_n^\pm$ from \eqref{eq:V}, this effect is more severe for the Lepski method.
\begin{figure}
\includegraphics[width=14cm]{img/quantilesDax}
\caption{Estimated generalized quantiles of negative jumps \textit{(left)} and positive jumps \textit{(right)} based on option prices from May 29, 2008, with four maturities $T$.}\label{fig:quantilesDax}
\end{figure}
Let us finally apply the estimation method to prices of DAX options from May 29, 2008. This data set\footnote{provided by the SFB 649 ``Economic Risk''} has already been studied by \cite{soehlTrabs2014}. Figure~\ref{fig:quantilesDax} shows the estimated quantiles $\tilde q_\tau^\pm$ for four different maturities between two and seven months and for $\tau\in\{0.2,0.4,\dots,4\}$. Due to this finer grid, the threshold value is set to $\eta=0.01$. As postulated by the stylized facts on financial data, negative jumps have a higher activity. Roughly, the intensity of small jumps is larger for short maturities while the tails are more heavy for longer maturities.
\section{Proofs}\label{sec:LevyProofs}
\subsection{Error analysis for Section~\ref{sec:quantileEst}}
To simplify the notation we will frequently use the definition $I_{h}:=[-\frac{1}{h},\frac{1}{h}]$. Note that the indicator function $\1_{\{|\phi_{\Delta,n}(u)|\ge(\Delta n)^{-1/2}\}}$ in the definition of $\hat\psi_n''$ equals one on $I_h$ with probability converging to one. This follows exactly from Lemma~5.1 by \cite{dattnerEtAl2013} and the bandwidth choices which we will consider.
\subsubsection{Drift and remainder }
All estimators are constructed based on the estimator $\hat\psi_n''(u)=\Delta^{-1}(\log(\phi_{\Delta,n}(u))''$ of $\psi''$ from \eqref{eq:psiPP}. In particular, $\hat\nu_h,\hat N_h$ and $\hat q_{\tau,h}$ only depend on the observations via $\hat\psi_n''$. As shown by \citet[Lem. 10]{nicklEtAl2013} the drift has no effect on the estimators:
\begin{lemma}
\label{lem:drift}Let $X_{k}:=Y_{k}-\Delta\gamma$ for $k=1,\dots,n.$
Then $X_{k}$ are distributed according to an infinitely divisible
distribution with characteristic triplet $(\sigma^{2},0,\nu)$. Denoting
the estimator $\hat\psi_n''$ based on $(X_{k})$ and $(Y_{k})$
by $\hat\psi''_{X,n}$ and $\hat{\psi}''_{Y,n}$, respectively, it holds $\hat\psi''_{X,n}=\hat\psi''_{Y,n}$
for all $u\in\R.$
\end{lemma}
\iffalse\begin{proof}
The drift causes a factor $e^{i\Delta\gamma u}$ in the empirical characteristic function $\phi_{\Delta,n,Y}$ such that
\[ \hat\psi_{n,Y}''(u)=\Delta^{-1} (\log(\phi_{\Delta,n,X}(u))+i\Delta\gamma u)''=\hat\psi_{n,X}''(u).
\]
\end{proof}\fi
Using
\begin{equation}
(\phi_{\Delta}^{-1})'=-\Delta\psi'\phi_{\Delta}^{-1},\qquad(\phi_{\Delta}^{-1})''=-\Delta\big(\psi''-\Delta(\psi'){}^{2}\big)\phi_{\Delta}^{-1},\label{eq:PhiDerivatives}
\end{equation}
the estimation error $\psi''-\hat{\psi}''_{n}=-\Delta^{-1}\log(\phi_{\Delta,n}/\phi_{\Delta})''$ can be linearized similarly to Proposition~21 in \citet{nicklEtAl2013}.
\begin{lemma}
\label{lem:Remainder}Let $\int x^{4+\delta}\nu(\d x)<\infty$ for
some $\delta>0$. For $h,\Delta\in(0,1)$ satisfying \\$n^{-1/2}(\log h^{-1})^{(1+\delta)/2}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}\to0$
as $n\to\infty$, it holds
\begin{align*}
& \sup_{|u|\le h^{-1}}\left|\hat{\psi}_{n}''(u)-\psi''(u)-\Delta^{-1}(\phi_{\Delta}^{-1}(\phi_{\Delta,n}-\phi_{\Delta}))''(u)\right|\\
= & \mathcal O_{P}\big((\Delta\|\psi'\big\|_{L^{\infty}(I_h)}+\Delta^{3/2}\|\psi'\big\|_{L^{\infty}(I_h)}^{2}+1)n^{-1}\Delta^{-1/2}\log(h^{-1})^{1+\delta}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}^{2}\big).
\end{align*}
\end{lemma}
\begin{proof}
Setting $F(y)=\log(1+y)$ and $\eta=(\phi_{\Delta,n}-\phi_{\Delta})/\phi_{\Delta}$,
we use $(F\circ\eta)''(u)=F'(\eta(u))\eta''(u)+F''(\eta(u))\eta'(u)^{2}$
to obtain
\begin{align}
|(F\circ\eta)''(u)-\eta''(u)| & \lesssim\|F''\|_{\infty}\Big(\eta(u)\eta''(u)+\eta'(u)^{2}\Big).\label{eq:MarkusTrick}
\end{align}
On the event $\Omega_{n}:=\{\sup_{|u|\le1/h}|(\phi_{\Delta,n}-\phi_{\Delta})(u)/\phi_{\Delta}(u)|\le1/2\}$
we thus obtain
\begin{align*}
& \sup_{\left|u\right|\le h^{-1}}\big|\log(\phi_{\Delta,n}/\phi_{\Delta})''(u)-(\phi_{\Delta}^{-1}(\phi_{\Delta,n}-\phi_{\Delta}))''(u)\big|\\
& \qquad\lesssim\|\eta\|_{L^{\infty}(I_h)}\|\eta''\|_{L^{\infty}(I_h)}+\|\eta'\|_{L^{\infty}(I_h)}^{2}.
\end{align*}
To estimate $\|\eta^{(k)}\|_{L^{\infty}(I_h)},k=0,1,2$, we use \eqref{eq:PhiDerivatives} and
$\left|\psi''(u)\right|\lesssim1$ to obtain
\begin{align*}
\sup_{u\in I_h}\left|(\phi_{\Delta}^{-1})'(u)\right| & \lesssim\Delta\|\psi'\big\|_{L^{\infty}(I_h)}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)},\\
\sup_{u\in I_h}\left|(\phi_{\Delta}^{-1})''(u)\right| & \lesssim(\Delta^{2}\|\psi'\big\|_{L^{\infty}(I_h)}^{2}+\Delta)\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}.
\end{align*}
Applying Theorem 1 by \citet{kappusReiss2010} and the moment assumption
on $\nu$, we obtain for $k=0,1,2$ and $\delta>0$
\[
\big\|(\phi_{\Delta,n}-\phi_{\Delta})^{(k)}\big\|_{L^{\infty}(I_h)}=\mathcal O_{P}(n^{-1/2}\Delta^{(k\wedge1)/2}(\log h^{-1})^{(1+\delta)/2}).
\]
This yields
\begin{align}
\|\eta\|_{L^{\infty}(I_h)} & =\mathcal O_{P}\big(n^{-1/2}(\log h^{-1})^{(1+\delta)/2}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}\big),\label{eq:supEta}\\
\|\eta'\|_{L^{\infty}(I_h)} & =\mathcal O_{P}\big((\Delta\|\psi'\big\|_{L^{\infty}(I_h)}+\Delta^{1/2})n^{-1/2}(\log h^{-1})^{(1+\delta)/2}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}\big),\nonumber \\
\|\eta''\|_{L^{\infty}(I_h)} & =\mathcal O_{P}\big((\Delta^{3/2}\|\psi'\big\|_{L^{\infty}(I_h)}+\Delta^{2}\|\psi'\big\|_{L^{\infty}(I_h)}^{2}+\Delta^{1/2})\nonumber \\
& \hspace{3em}\times n^{-1/2}(\log h^{-1})^{(1+\delta)/2}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}\big).\nonumber
\end{align}
The bound (\ref{eq:supEta}) and $n^{-1/2}(\log h^{-1})^{(1+\delta)/2}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}\to0$,
yield $P(\Omega_{n})\to1$ which implies
the assertion.
\end{proof}
\subsubsection{Convergence rates for distribution function estimation}
Because it is a bit more difficult to derive the convergence rates for the distribution function estimator than for the density estimator, we study this problem first. The corresponding results for $\hat\nu_n$
can the be proved analogously.
We decompose the estimation error of the distribution function estimator
$\hat{N}_{h}$ into
\begin{align}
\hat{N}_{h}(t)-N(t)= & \int g_{t}(x)\big(K_{h}\ast\big(y^{2}\nu\big)-x^{2}\nu\big)(\d x)\nonumber \\
& +\int g_{t}(x)\F^{-1}\Big[\Big(\psi''(u)-\hat{\psi}''_{n}(u)\Big)\F K(hu)\Big](x)\d x+\sigma^{2}\int g_{t}(x)K_{h}(x)\d x\nonumber \\
=: & B_{n}(t)+S_{n}(t)+V_{n}(t),\label{eq:Decomp}
\end{align}
where $B_{n}$ is the deterministic error term, $S_{n}$ is the stochastic
error term and $V_{n}$ is the error due to the unknown volatility
$\sigma^{2}$. The error term $V_{n}$ is negligible:
\begin{lemma}
\label{lem:vola}Grant Assumption (\ref{eq:propKernel}) on the kernel
with order $p\ge s+1$. Then $V_{n}(t),t\neq0,$ as defined in (\ref{eq:Decomp})
satisfies $|V_{n}(t)|\lesssim\sigma^{2}|t|^{-s-3}h^{s+1}.$\end{lemma}
\begin{proof}
We estimate
\begin{align*}
|V_{n}|\le\sigma^{2}\int|g_{t}(x)K_{h}(x)|\d x\le & \sigma^{2}\|x^{-s-1}g_{t}(x)\|_{\infty}\|x^{s+1}K_{h}(x)\|_{L^{1}}\\
= & \sigma^{2}|t|^{-s-3}h^{s+1}\|x^{s+1}K(x)\|_{L^{1}}.\qedhere
\end{align*}
\end{proof}
For the bias we apply the following:
\begin{prop}\label{prop:bias}
Suppose $\|x^{2}\nu\|_{L^{1}}<\infty$ and let $U\subset\R$
be an open set. If $\nu$ admits a Lebesgue
density on $U$ in $C^{s}(U)$ for some $s>-1$ and if the kernel satisfies
(\ref{eq:propKernel}) with order $p\ge s+1$, then
\[
|B_{n}(t)|=\Big|\int g_{t}(x)\big(K_{h}\ast\big(y^{2}\nu(\d y)\big)-x^{2}\nu\big)(\d x)\Big|\lesssim(|t|^{-s-4}\vee1)h^{s+1},\quad\text{for all}\quad t\in U.
\]
\end{prop}
\begin{proof}
Without loss of generality let $t<0.$ Using Fubini's theorem, we
rewrite
\begin{align}
\int_{\R}g_{t}(x)\big(K_{h}\ast\big(y^{2}\nu(\d y)\big)-x^{2}\nu\big)(\d x)=K_{h}\ast g_{t}(-\bull)\ast(x^{2}\nu)(0)-g_{t}(-\bull)\ast(x^{2}\nu)(0).\label{eqRewrBias}
\end{align}
Denoting $\overline{N}(t)=\int_{-\infty}^{t}\nu(\d x),t<0,$ integration
by parts yields
\begin{align*}
g_{t}(-\bull)\ast(x^{2}\nu)(y) & =\int_{-\infty}^{t+y}\frac{x^{2}}{(x-y)^{2}}\nu(\d x)\\
& =\frac{(t+y)^2}{t^2}\overline{N}(t+y)+\int_{-\infty}^{0}\frac{2(x+t+y)y}{(x+t)^{3}}\overline{N}(x+t+y)\d x\\
& =\frac{(t+y)^2}{t^2}\overline{N}(t+y)+2y\left(((\bull+t)^{-2}\1_{(-\infty,0]})\ast\overline{N}(t-\bull)\right)(-y) \\
&\qquad+2y^2\left(((\bull+t)^{-3}\1_{(-\infty,0]})\ast\overline{N}(t-\bull)\right)(-y).
\end{align*}
From this representation we see for some sufficiently small $\delta\in(0,|t|)$
that $g_{t}(-\bull)\ast(x^{2}\nu)\in C^{s+1}((-\delta/2,\delta/2))$
owing to $\overline{N}(t+\bull)\in C^{s+1}((-\delta,\delta))$ and $(\bull+t)^{-3}\1_{(-\infty,0]}\in C^{s+1}((-\delta/2,\delta/2)^{c})$.
The corresponding H\"older norm of the latter is of order $t^{-s-4}$.
With a standard Taylor expansion argument and applying the order of
the kernel, the approximation error is of the order $(|t|^{-s-4}\vee1)h^{s+1}$ by
(\ref{eqRewrBias}).
\end{proof}
Lemma~\ref{lem:Remainder} motivates the following definition of
the linearized stochastic error term
\begin{equation}
L_{\Delta,n}(t):=-\Delta^{-1}\int g_{t}(x){\cal F}^{-1}[\F K(h\bull)\big(\phi_{\Delta}^{-1}(\phi_{\Delta,n}-\phi_{\Delta})\big)''](x)\d x.\label{eq:linError}
\end{equation}
Using \eqref{eq:PhiDerivatives} and defining the regularized Fourier multiplier
\[
m_{\Delta,h}:=\frac{\F K(h\bull)}{\phi_{\Delta}},
\]
we decompose the linearized stochastic error further into
\begin{align}
L_{\Delta,n}(t)= & -\frac{1}{\Delta}\int g_{t}(x){\cal F}^{-1}[\F K(h\bull)\big(\phi_{\Delta}^{-1}(\phi_{\Delta,n}-\phi_{\Delta})''+2(\phi_{\Delta}^{-1})'(\phi_{\Delta,n}-\phi_{\Delta})'\nonumber \\
& \quad+(\phi_{\Delta}^{-1})''(\phi_{\Delta,n}-\phi_{\Delta})\big)](x)\d x\nonumber \\
= & \underbrace{-\frac{1}{\Delta}\int g_{t}(x){\cal F}^{-1}[m_{\Delta,h}(\phi_{\Delta,n}''-\phi_{\Delta}'')](x)\d x}_{=:M_{\Delta,n}(t)}\notag\\
&\qquad+2\int g_{t}(x){\cal F}^{-1}[m_{\Delta,h}\psi'(\phi_{\Delta,n}'-\phi_{\Delta}')](x)\d x\nonumber \\
& \qquad+\int g_{t}(x){\cal F}^{-1}[m_{\Delta,h}\big(\psi''-\Delta(\psi'){}^{2}\big)(\phi_{\Delta,n}-\phi_{\Delta})](x)\d x.\label{eq:DecompLin}
\end{align}
In the following, we will refer to $M_{\Delta,n}$ as the main stochastic
error term.
\begin{prop}\label{prop:variance}
Let $U\subset\R$ be an open set, $t\in U$
and $\alpha,\beta,s,r,R>0,m>4$. Let the kernel satisfy (\ref{eq:propKernel})
with $p\ge1$. Then we obtain for $h,\Delta\in(0,1)$ and $\eta\in(0,1)$
\begin{align*}
\sup_{(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)}\E\Big[\sup_{|t|>\eta}|L_{\Delta,n}-M_{\Delta,n}|(t)\Big] & \lesssim\eta^{-2}n{}^{-1/2}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)},\\
\sup_{(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)}\E\Big[\sup_{|t|>\eta}|L_{\Delta,n}-M_{\Delta,n}|(t)\Big] & \lesssim\frac{\log(h^{-1})+\Delta^{1/2}h^{-1}+\Delta h^{-2}}{\eta^{2}n^{1/2}}e^{r\Delta h^{-\beta}}.
\end{align*}
Moreover,
\[
\sup_{(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)}\E\Big[\sup_{|t|>\eta}|M_{\Delta,n}(t)|\Big]\lesssim\eta^{-2}(n\Delta)^{-1/2}\log(h^{-1})e^{r\Delta h^{-\beta}}.
\]
\end{prop}
\begin{remark}
Note that
\begin{align*}
\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)} & \lesssim\begin{cases}
h^{-\Delta\alpha+1/2} & \text{for }\Delta\alpha>1/2,\\
(\log h^{-1})^{1/2} & \text{for }\Delta\alpha=1/2,\\
1 & \text{for }\Delta\alpha\in(0,1/2).
\end{cases}
\end{align*}
\end{remark}
\begin{proof}
Due to $\|g_{t}\|_{L^{1}}=|t|^{-1},\|g_{t}\|_{BV}\le2t^{-2}$, we
obtain $|\F g_{t}(u)|\lesssim(t^{-1}\vee t^{-2})(1+|u|)^{-1},u\in\R$.
Using Plancherel's identity, we estimate for $|t|>\eta$
\begin{align*}
|L_{\Delta,n}-M_{\Delta,n}|(t)\le & \pi^{-1}\Big|\int\F g_{t}(-u)m_{\Delta,h}(u)\psi'(u)(\phi_{\Delta,n}-\phi_{\Delta})'(u)\d u\Big|\\
& +(2\pi)^{-1}\Big|\int\F g_{t}(-u)m_{\Delta,h}(u)\big(\psi''(u)-\Delta\psi'(u){}^{2}\big)(\phi_{\Delta,n}-\phi_{\Delta})(u)\d u\Big|\\
\lesssim & \eta^{-2}\int(1+|u|)^{-1}\big|m_{\Delta,h}(u)\psi'(u)(\phi_{\Delta,n}-\phi_{\Delta})'(u)\big|\d u\\
& +\eta^{-2}\int(1+|u|)^{-1}\big|m_{\Delta,h}(u)\psi''(u)(\phi_{\Delta,n}-\phi_{\Delta})(u)\big|\d u\\
& +\eta^{-2}\Delta\int(1+|u|)^{-1}\big|m_{\Delta,h}(u)\psi'(u){}^{2}(\phi_{\Delta,n}-\phi_{\Delta})(u)\big|\d u.
\end{align*}
Due to $\E[Y_{1}^{2l}]\lesssim\Delta^{l\wedge1},l=0,1,2$, we have, moreover,
\begin{equation}
\sup_{u\in\R}\E[(\phi_{\Delta,n}^{(l)}-\phi_{\Delta}^{(l)})^{2}(u)]\lesssim n^{-1}\Delta^{(l\wedge1)},\quad\text{for }l=0,1,2.\label{eq:phiMom}
\end{equation}
Fubini's theorem and Jensen's inequality yield
\begin{align}
& \E\Big[\sup_{|t|>\eta}|L_{\Delta,n}-M_{\Delta,n}|(t)\Big]\nonumber \\
\lesssim & \eta^{-2}\Big(\int(1+|u|)^{-1}\big|m_{\Delta,h}(u)\psi'(u)\big|\E\big[(\phi'_{\Delta,n}-\phi'_{\Delta})^{2}(u)\big]^{1/2}\d u\nonumber \\
& \qquad+\int(1+|u|)^{-1}\big|m_{\Delta,h}(u)\psi''(u)|\E[(\phi_{\Delta,n}-\phi_{\Delta})^{2}(u)]^{1/2}\d u\nonumber \\
& \qquad+\Delta\int(1+|u|)^{-1}\big|m_{\Delta,h}(u)\psi'(u){}^{2}|\E[(\phi_{\Delta,n}-\phi_{\Delta})^{2}(u)]^{1/2}\d u\Big)\nonumber \\
\lesssim & n^{-1/2}\eta^{-2}\Big(\Delta^{1/2}\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\psi'(u)\big\|_{L^{1}}+\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\psi''(u)\big\|_{L^{1}}\nonumber \\
& \qquad+\Delta\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\psi'(u){}^{2}\big\|_{L^{1}}\Big).\label{eq:L-M}
\end{align}
To deal with $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$,
we note that the assumptions $\|x^{4}\nu\|_{L^{1}}<\infty$ and $\|x\nu\|_{\infty}<\infty$
imply $x\nu,x^{2}\nu\in L^{1}(\R)\cap L^{2}(\R)$. Moreover, $\sigma^{2}=0$
and by Lemma~\ref{lem:drift} we can assume $\gamma_0=\gamma-\int x\d\nu=0$ such that $i\psi'=-\F[x\nu]$ and $\psi''=-\F[x^{2}\nu]$. Therefore,
we estimate (\ref{eq:L-M}) with use of the Cauchy--Schwarz inequality
\begin{align}
& \E\big[\sup_{|t|>\eta}|L_{\Delta,n}-M_{\Delta,n}|(t)\big]\nonumber \\
\lesssim & n^{-1/2}\eta^{-2}\Big(\Delta^{1/2}\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\|_{L^2}\|\psi'\big\|_{L^{2}}+\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\|_{L^{2}}\|\psi''\big\|_{L^{2}}\nonumber \\
& \qquad+\Delta\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\|_{L^{2}}\|\psi'\big\|_{L^{2}}\|\psi'\|_{\infty}\Big)\nonumber \\
\lesssim & n^{-1/2}\eta^{-2}\big(\Delta^{1/2}\|x\nu\|_{L^{2}}+\|x^{2}\nu\|_{L^{2}}+\Delta\|x\nu\|_{L^{2}}\|x\nu\|_{L^{1}}\big)\big\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)},\label{eq:nonCritMild}
\end{align}
which yields the assertion for $\mathcal{D}^{s}(\alpha,m,U,R)$.
Now we consider the case $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)$.
The exponential decay of $\phi_{\Delta}$, the properties of $K$,
$|\psi'(u)|\lesssim1+|u|$ and $|\psi''(u)|\lesssim1$ yield
\begin{align*}
\Delta^{1/2}\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\psi'(u)\big\|_{L^{1}}\lesssim & \Delta^{1/2}\big\| m_{\Delta,h}\big\|_{L^{1}}\\
\lesssim&\Delta^{1/2}\int_{-1/h}^{1/h}e^{r\Delta|u|^{\beta}}\d u\lesssim\Delta^{1/2}h^{-1}\exp\big(r\Delta h^{-\beta}\big),\\
\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\psi''(u)\|_{L^{1}}\lesssim & \big\|(1+u)^{-1}m_{\Delta,h}(u)\big\|_{L^{1}}\lesssim\log(h^{-1})\exp\big(r\Delta h^{-\beta}\big),\\
\Delta\big\|(1+|u|)^{-1}m_{\Delta,h}(u)\psi'(u){}^{2}\big]\big\|_{L^{1}}\lesssim & \Delta\big\|(1+u)m_{\Delta,h}(u)\big\|_{L^{1}}\lesssim\Delta h^{-2}\exp\big(r\Delta h^{-\beta}\big).
\end{align*}
We conclude the claimed estimate for $|L_{\Delta,n}-M_{\Delta,n}|$ in $\mathcal{E}^{s}(\beta,m,U,r,R)$
by plugging these estimates into (\ref{eq:L-M}). Similarly, we estimate
\begin{align*}
\E\Big[\sup_{|t|>\eta}|M_{\Delta,n}(t)|\Big] & \lesssim\eta^{-2}\Delta^{-1}\E\Big[\int(1+|u|)^{-1}m_{\Delta,h}(u)(\phi_{\Delta,n}''-\phi_{\Delta}'')(u)\d u\Big]\\
& =\eta^{-2}(n\Delta)^{-1/2}\|(1+|u|)^{-1}m_{\Delta,h}(u)\|_{L^{1}}\\
&\lesssim\eta^{-2}(n\Delta)^{-1/2}\log(h^{-1})\exp\big(r\Delta h^{-\beta}\big).\qedhere
\end{align*}
\end{proof}
For the main stochastic error term in the mildly ill-posed case we
will need the following concentration result:
\begin{prop}
\label{prop:mainStochErr}Let $U\subset\R$ be an open set, $t\in U$
and $\alpha,s,R>0,m>4$ and let the kernel satisfy (\ref{eq:propKernel})
with $p\ge1$. Then there is some $c>0$ such that for any $h,\Delta,\eta\in(0,1)$
and $\kappa_{0}>0$
\begin{align*}
\sup_{(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)}\sup_{|t|>\eta} & P\Big(|M_{\Delta,n}(t)|>\kappa_{0}(\eta^{-1}\vee1)(n\Delta)^{-1/2}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}\Big)\\
& \hspace{6em}\le2\exp\Big(-\frac{c\kappa_{0}^{2}}{1+\kappa_{0}(\eta^{-1}\vee1)(hn\Delta)^{-1/2}}\Big).
\end{align*}
\end{prop}
\begin{proof}
We represent $M_{\Delta,n}(t)$ as sum of i.i.d. random variables
via
\[
M_{\Delta,n}(t)=\sum_{k=1}^{n}(\xi_{k}(t)-\E[\xi_{k}(t)]),\quad\xi_{k}(t):=(n\Delta)^{-1}\int g_{t}(x)\F^{-1}\big[m_{\Delta,h}(u)Y_{k}^{2}e^{iuY_{k}}\big](x)\d x.
\]
Applying Plancherel's identity, $\xi_{k}(t)$ can be rewritten as
\begin{align*}
\xi_{k}(t) & =\frac{1}{2\pi n\Delta}Y_{k}^{2}\int\F g_{t}(-u)m_{\Delta,h}(u)e^{iuY_{k}}\d u=\frac{1}{n\Delta}Y_{k}^{2}\F^{-1}[\F g_{t}(-\bull)m_{\Delta,h}](-Y_{k}).
\end{align*}
To estimate $\Var(\xi_{k})\le\E[\xi_{k}^{2}]$, we use $g_{t}\in BV(\R)$
to decompose $g_{t}=g_{t}^{s}+g_{t}^{c}$ into a singular component
and a continuous component satisfying for $r>\Delta\alpha$
\begin{align*}
\max\left\{ \F g_{t}^{s}(u),\F[xg_{t}^{s}](u),\F[x^{2}g_{t}^{s}](u)\right\} &\lesssim(t^{-2}\vee1)(1+|u|)^{-1},\\
\max\Big\{\sup_{|t|>\eta}\|g_{t}^c\|_{C^r},\sup_{|t|>\eta}\|x^2g_{t}^c\|_{C^r}\Big\}&\lesssim1.
\end{align*}
This allows to decompose
\begin{align}
\E[\xi_{k}^{2}] & \le\frac{2}{(n\Delta)^{2}}\Big(\E\Big[Y_{1}^{4}\F^{-1}\big[\F g_{t}^{s}(-\bull)m_{\Delta,h}\big](-Y_{1})^{2}\Big]\notag\\
&\qquad\qquad+\E\Big[Y_{1}^{4}\F^{-1}\big[\F g_{t}^{c}(-\bull)m_{\Delta,h}\big](-Y_{1})^{2}\Big]\Big) =:\frac{2}{(n\Delta)^{2}}(E_{s}+E_{c}).\label{eq:VarM}
\end{align}
To estimate $E_{c}$ in (\ref{eq:VarM}), we apply the Fourier multiplier
Theorem~5 in \cite{trabs2014} to see that
\[
\|\F^{-1}\big[\F g_{t}^{c}(-\bull)m_{\Delta,h}\big]\|_{\infty}\lesssim\|g_{t}^{c}\ast K_{h}\|_{C^{r}}\le\|K\|_{L^{1}}\|g_{t}^{c}\|_{C^{r}}
\]
for any $r>\Delta\alpha$. Consequently, $(n\Delta)^{-2}E_{s}\lesssim(n\Delta)^{-2}\E[Y_{1}^{4}]\lesssim n^{-2}\Delta^{-1}$
because $Y_{1}$ has finite fourth moments due to $\|x^{4}\nu\|_{L^{1}}<\infty$.
It remains to bound $E_{s}$ from (\ref{eq:VarM}). Using again $\gamma_0=0$ by Lemma~\ref{lem:drift}, we infer from
$\F[ixP_{\Delta}]=\phi_{\Delta}'=\Delta\psi'\phi=\Delta\F[ix\nu]\phi_{\Delta}$
that
\begin{equation}
xP_{\Delta}=\Delta(x\nu)\ast P_{\Delta}\label{eq:xP}
\end{equation}
and thus $xP_{\Delta}$ has a bounded density satisfying $\|xP_{\Delta}\|_{\infty}\le\Delta\|x\nu\|_{\infty}$.
Together with the Cauchy--Schwarz inequality and Plancherel's identity
we obtain
\begin{align}
E_{s}\le & \Delta\|x\nu\|_{\infty}\int|y|^{3}\Big(\F^{-1}\big[\F g_{t}^{s}(-\bull)m_{\Delta,h}\big](-y)\Big)^{2}\d y\nonumber \\
= & \Delta\|x\nu\|_{\infty}\int\Big|\F^{-1}\big[(\F g_{t}^{s}(-\bull)m_{\Delta,h})''\big](-y)\F^{-1}\big[(\F g_{t}^{s}(-\bull)m_{\Delta,h})'\big](-y)\Big|\d y\nonumber \\
\le & \Delta\|x\nu\|_{\infty}\left\Vert (\F g_{t}^{s}(-\bull)m_{\Delta,h})''\right\Vert _{L^{2}}\left\Vert (\F g_{t}^{s}(-\bull)m_{\Delta,h})'\right\Vert _{L^{2}}.\label{eq:varEs}
\end{align}
The derivatives of the regularized Fourier multiplier are given by
\begin{align}
m_{\Delta,h}'(u) & =ih\frac{\F[xK](hu)}{\phi_{\Delta}(u)}-\Delta\psi'(u)m_{\Delta,h}(u)=ih\frac{\F[xK](hu)}{\phi_{\Delta}(u)}-i\Delta\F[x\nu](u)m_{\Delta,h}(u),\label{eq:mP}\\
m_{\Delta,h}''(u) & =-h^{2}\frac{\F[x^{2}K](hu)}{\phi_{\Delta}(u)}-2i\Delta h\psi'(u)\frac{\F[xK](hu)}{\phi_{\Delta}(u)}-\Delta\big(\psi''(u)(u)+\Delta\psi'(u)^{2}\big)m_{\Delta,h}(u).\label{eq:mPP}
\end{align}
To bound the $L^{2}$-norms in \eqref{eq:varEs}, we use the properties
of $K$, the decay assumption on $\phi_{\Delta}$ and $\psi',\psi''\in L^\infty(\R)$ to obtain
\begin{align}
\|(\F g_{t}^{s}(-\bull)m_{\Delta,h})'\|_{L^{2}}\le & \|\F[xg_{t}^{s}](-\bull)m_{\Delta,h}\|_{L^{2}}+\|\F[g_{t}^{s}](-\bull)m_{\Delta,h}'\|_{L^{2}}\nonumber \\
\lesssim & (t^{-2}\vee1)(1+h+\Delta)\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)},\nonumber \\
\|(\F g_{t}^{s}(-\bull)m_{\Delta,h})''\|_{L^{2}}\le & \|\F[x^{2}g_{t}^{s}](-\bull)m_{\Delta,h}\|_{L^{2}} +2\|\F[xg_{t}^{s}](-\bull)m_{\Delta,h}'\|_{L^{2}}+\|\F g_{t}^{s}(-\bull)m_{\Delta,h}''\|_{L^{2}}\notag\\
\lesssim & (t^{-2}\vee1)(1+h+\Delta+h^{2}+\Delta h)\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}.\label{eq:L2gm}
\end{align}
Therefore, $(\Delta n)^{-2}E_{s}\lesssim (t^{-2}\vee1)n^{-2}\Delta^{-1}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}^{2}$ which implies
\begin{equation}
\Var(\xi_{k}(t))\lesssim\frac{t^{-2}\vee1}{n^{2}\Delta}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}^{2}.\label{eq:VarXi}
\end{equation}
Using (\ref{eq:mP}), (\ref{eq:mPP}), $x\nu\in L^{2}(\R)$, $|\F g_{t}(u)|\lesssim(t^{-2}\vee1)(1+|u|)^{-1}$
and $\|xg_{t}\|_{L^{2}}\lesssim|t|^{-1}$, we deterministically bound
$\xi_{k}(t)$ by
\begin{align*}
|\xi_{k}(t)|\le & (n\Delta)^{-1}\big\|\F^{-1}[(\F g_{t}(-\bull)m_{\Delta,h})'']\big\|_{\infty}\allowdisplaybreaks\\
\le & (n\Delta)^{-1}\Big(\big\|\F^{-1}\big[\F[x^{2}g_{t}](-\bull)m_{\Delta,h}\big]\big\|_{\infty}+2\|\F[xg_{t}](-\bull)m_{\Delta,h}'\|_{L^{1}}+\|\F g_{t}(-\bull)m_{\Delta,h}''\|_{L^{1}}\Big)\allowdisplaybreaks\\
\lesssim & (n\Delta)^{-1}\Big(\big\|\F^{-1}\big[\F[x^{2}g_{t}](-\bull)m_{\Delta,h}\big]\big\|_{\infty}+\|xg_{t}\|_{L^{2}}\|m_{\Delta,h}'\|_{L^{2}}\\
& +\|(1+|u|)\F g_{t}(u)\|_{\infty}\|(1+|u|)^{-1}m_{\Delta,h}''(u)\|_{L^{1}}\Big)\allowdisplaybreaks\\
\lesssim & (n\Delta)^{-1}\Big(\|\F[x^{2}g_{t}^{s}](-\bull)m_{\Delta,h}\|_{L^{1}}+\|\F^{-1}\big[\F[x^{2}g_{t}^{c}](-\bull)m_{\Delta,h}\big]\|_{\infty}\\
& +\|xg_{t}\|_{L^{2}}\big(h\|(1+|u|)^{\Delta\alpha}\|_{L^{2}(I_h)}+\Delta \|x\nu\|_{L^{1}}\|(1+|u|)^{\Delta\alpha}\|_{L^{2}(I_h)}\big)\\
& +(t^{-2}\vee1)(\Delta+h^{2}+\Delta h)\|(1+|u|)^{\Delta\alpha-1}\|_{L^{1}(I_h)}\Big)\allowdisplaybreaks\\
\lesssim & (n\Delta)^{-1}\|\F[x^{2}g_{t}^{s}](-\bull)m_{\Delta,h}\|_{L^{1}}+(n\Delta)^{-1}\|\F^{-1}\big[\F[x^{2}g_{t}^{c}](-\bull)m_{\Delta,h}\big]\|_{\infty}\\
& +(n\Delta)^{-1}(t^{-2}\vee1)h^{-1/2}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}.
\end{align*}
The term corresponding to the singular part $g_{t}^{s}$ in the previous bound can be estimated by
\begin{align*}
(n\Delta)^{-1}\big\|\F[x^{2}g_{t}^{s}](-\bull)m_{\Delta,h}\big\|_{L^{1}} & \lesssim(n\Delta)^{-1}\|(1+|u|)\F[x^{2}g_{t}^{s}](u)\|_{\infty}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{1}(I_h)}\\
& \lesssim(n\Delta)^{-1}(t^{-2}\vee1)h^{-1/2}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}.
\end{align*}
For the continuous part $g_{t}^{c}$ we apply the Fourier multiplier
theorem as above to see that $$\|\F^{-1}\big[\F[x^{2}g_{t}^{c}](-\bull)m_{\Delta,h}\big]\|_{\infty}\lesssim\|x^{2}g_{t}^{c}\|_{C^{r}}$$
for any $r>\Delta\alpha$. Therefore,
\begin{align}
|\xi_{k}(t)|\le & (n\Delta)^{-1}\big\|\F^{-1}[(\F g_{t}(-\bull)m_{\Delta,h})'']\big\|_{\infty}\nonumber \\
\lesssim & (n\Delta)^{-1}(t^{-2}\vee1)h^{-1/2}\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}.\label{eq:DetXi}
\end{align}
Using (\ref{eq:VarXi}) and (\ref{eq:DetXi}), Bernstein's inequality
yields for some constant $c>0$ the claimed concentration result.
\end{proof}
Combining the previous results, we obtain minimax convergence rates
for estimating the (generalized) distribution function of the jump
measure.
\begin{proof}[Proof of Proposition \ref{prop:DistFunct}]
In the following, $t$ is fixed and thus omitted in the constants.
Using the error decomposition (\ref{eq:Decomp}), Lemma~\ref{lem:vola},
Proposition~\ref{prop:bias}, we obtain
\begin{align*}
|\hat{N}_{h}(t)-N(t)|\le & |B_{n}(t)|+|S_{n}(t)|+|V_{n}(t)|\\
\lesssim & h^{s+1}+|S_{n}(t)|.
\end{align*}
Using $|\F g_{t}(u)|\lesssim(1+|u|)^{-1}$, Plancherel's identity
and Lemma~\ref{lem:Remainder} yield for the stochastic error from
(\ref{eq:Decomp}) and the linearized stochastic error term $L_{\Delta,n}$
defined in (\ref{eq:linError})
\begin{align}
& \big|S_{n}(t)-L_{\Delta,n}(t)\big|\nonumber \\
= & \left|\int g_{t}(x)\F^{-1}\Big[\F K(hu)\Big(\hat{\psi}''_{n}(u)-\psi''(u)-\Delta^{-1}\big(\phi_{\Delta}^{-1}(\phi_{\Delta,n}-\phi_{\Delta})\big)''(u)\Big)\Big](x)\d x\right|\nonumber \\
\lesssim & \sup_{\left|u\right|\le h^{-1}}\big|\hat{\psi}_{n}''(u)-\psi''(u)-\Delta^{-1}\big(\phi_{\Delta}^{-1}(\phi_{\Delta,n}-\phi_{\Delta})\big)''(u)\big|\int(1+|u|)^{-1}|\F K(hu)|\d u\nonumber \\
\le & \sup_{\left|u\right|\le h^{-1}}\big|\hat{\psi}_{n}''(u)-\psi''(u)-\Delta^{-1}\big(\phi_{\Delta}^{-1}(\phi_{\Delta,n}-\phi_{\Delta})\big)''(u)\big|\|K\|_{L^{1}}\int_{-1/h}^{1/h}(1+|u|)^{-1}\d u\nonumber \\
= & \mathcal{O}_{P}\big((\Delta\|\psi'\big\|_{L^{\infty}(I_h)}+\Delta^{3/2}\|\psi'\big\|_{L^{\infty}(I_h)}^{2}+1)\frac{|\log h|^{2+\delta}}{n\Delta^{1/2}}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}^{2}\big),\label{eq:linearization}
\end{align}
provided that $n^{-1/2}(\log h^{-1})^{(1+\delta)/2}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}\to0$.
The latter condition is satisfied for the choices $h=h_{n,\Delta}$ in
both cases.
Let $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$. We conclude
from (\ref{eq:linearization}), where $\psi'$ is uniformly bounded
and $\phi_{\Delta}$ decays polynomially, and Propositions~\ref{prop:variance}
and \ref{prop:mainStochErr} for $\Delta\alpha>1/2$
\begin{align*}
S_{n}(t)= & L_{\Delta,n}(t)+\mathcal{O}_{P}\big(n^{-1}\Delta^{-1/2}\log(h^{-1})^{2+\delta}h^{-2\Delta\alpha}\big)\\
= & \mathcal{O}_{P}\big((n\Delta)^{-1/2}h^{-\Delta\alpha+1/2}+n^{-1}\Delta^{-1/2}\log(h^{-1})^{2+\delta}h^{-2\Delta\alpha}\big)\\
= & \mathcal{O}_{P}\big((n\Delta)^{-1/2}h^{-\Delta\alpha+1/2}(1+n^{-1/2}\log(h^{-1})^{2+\delta}h^{-\Delta\alpha-1/2})\big).
\end{align*}
Therefore, we obtain the rate $r_{n,\Delta}$ by plugging in $h=h_{n,\Delta}=(n\Delta)^{-1/(2s+2\Delta\alpha+1)}$ and similarly for $\Delta\alpha\le 1/2$.
Let us consider the case $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)$.
Owing to $\|\psi'\big\|_{L^{\infty}(I_h)}\lesssim h^{-1}$ and
the exponential decay of $\phi_{\Delta}$, we obtain from (\ref{eq:linearization})
and Proposition~\ref{prop:variance} that
\begin{align*}
S_{n}(t)= & \mathcal{O}_{P}\big((n\Delta)^{-1/2}\big(\log(h^{-1})+\Delta h^{-1}+\Delta^{3/2}h^{-2}\\
& \qquad+n^{-1/2}\log(h^{-1})^{2+\delta}(\Delta h^{-1}+\Delta^{3/2}h^{-2}+1)\big)\exp(r\Delta h^{-\beta})\big).
\end{align*}
Therefore, plugging in $h=h_{n,\Delta}=(\frac{\delta}{2r}\frac{\log(n\Delta)}{\Delta})^{-1/\beta},\delta\in(0,1)$,
yields
\begin{align*}
&\big|\hat{N}_{h}(t)-N(t)\big|\\
&\quad=\mathcal{O}_{P}\Big(\Big(\frac{\log(n\Delta)}{\Delta}\Big)^{-(s+1)/\beta}+(n\Delta)^{-(1-\delta)/2}\log(n\Delta)^{2/\beta}(|\log\Delta|+\Delta^{1-1/\beta}+\Delta^{3/2-2/\beta})\Big).
\end{align*}
\end{proof}
\subsubsection{Uniform loss for density estimation}
Applying a similar decomposition as in (\ref{eq:Decomp}) and the linearization
Lemma~\ref{lem:Remainder}, we obtain
\begin{align}
&\hat{\nu}_{h}(t)-\nu(t)\notag\\
=&\frac{1}{t^2}\left(\big(K_{h}\ast\big(y^{2}\nu\big)-x^{2}\nu\big)(t)+\F^{-1}\Big[\Big(\psi''(u)-\hat{\psi}''_{n}(u)\Big)\F K(hu)\Big](t)+\sigma^{2}K_{h}(t)\right)\label{eq:DecompDens}\\
= & \frac{1}{t^2}\Big(\underbrace{(K_{h}\ast(y^{2}\nu)-x^{2}\nu)(t)}_{=:B^{\nu}(t)}-\underbrace{\frac{1}{\Delta}\F^{-1}\Big[\F K(hu)\Big(\frac{\phi_{\Delta,n}-\phi_{\Delta}}{\phi_{\Delta}}\Big)''(u)\Big](t)}_{=:L^\nu_{\Delta,n}(t)}+R_{\Delta,n}+\sigma^{2}K_{h}(t)\Big)\notag
\end{align}
for some remainder $R_{\Delta,n}$ which is of order
\begin{align}
|R_{\Delta,n}| & =\mathcal{O}_{P}\big((\Delta\|\psi'\big\|_{L^{\infty}(I_h)}+\Delta^{3/2}\|\psi'\big\|_{L^{\infty}(I_h)}^{2}+1)\label{eq:RemainderDens}\\
& \qquad\qquad\times n^{-1}\Delta^{-1/2}h^{-1}\log(h^{-1})^{1+\delta}\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}^{2}\big).\nonumber
\end{align}
We will need the following concentration result for the main stochastic
error term of the density estimation problem
\begin{equation}
M^\nu_{\Delta,n}(t)=-\frac{1}{n\Delta}\sum_{k=1}^{n}\F^{-1}\Big[m_{\Delta,h}\big(Y_{k}^{2}e^{iuY_{k}}-\E[Y_{k}^{2}e^{iuY_{k}}]\big)\Big](t),\label{eq:MainStochErr}
\end{equation}
where we recall $m_{\Delta,h}=\F K(h\bull)/\phi_{\Delta}$. We will prove it analogously to Proposition~\ref{prop:mainStochErr}.
\begin{lemma}\label{lem:concMain}
Let $\alpha,R>0,m>4,U\subset\R$ and $t\neq0$ and let the kernel satisfy \eqref{eq:propKernel} for $p\ge1$. If $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$, then there is some constant $c>0$, depending only on $\alpha,R$,
such that for any $\kappa_{0}>0$ and any $n\in\mathbb{N},\Delta,h>0$
\[
P\left(|M^\nu_{\Delta,n}(t)|>\kappa_{0}(\Delta n)^{-1/2}h^{-\Delta\alpha-1/2}\right)\le2\exp\left(-\frac{c\kappa_{0}^{2}}{(1+t^{3})(1+\kappa_{0}(n\Delta h)^{-1/2})}\right).
\]
\end{lemma}
\begin{proof}
We apply Bernstein's inequality to the sum of the independent and
centered random variables
\[
M^\nu_{\Delta,n}(t)=-\sum_{k=1}^{n}\big(\xi_{k}-\E[\xi_{k}]\big),\quad\text{with}\quad\xi_{k}:=\frac{1}{n\Delta}\F^{-1}\Big[m_{\Delta,h}(u)Y_{k}^{2}e^{iuY_{k}}\Big](t).
\]
$\Var(\xi_{k})$ can be estimated similarly to (\ref{eq:varEs}).
We obtain by (\ref{eq:xP}), the Cauchy--Schwarz inequality, Plancherel's
identity, \eqref{eq:mP} and \eqref{eq:mPP} that
\begin{align*}
\Var(\xi_{k})\le&\E\big[\xi_{k}^{2}\big]= (n\Delta)^{-2}\E\Big[Y_{k}^{4}\Big(\F^{-1}\Big[m_{\Delta,h}(u)e^{-iut}\Big](-Y_{k})\Big)^{2}\Big]\\
\le & \frac{1}{n^{2}\Delta}\|x\nu\|_{\infty}\Big\|y^{2}\F^{-1}\Big[m_{\Delta,h}(u)e^{-iut}\Big](-y)\Big\|_{L^{2}}\Big\|y\F^{-1}\Big[m_{\Delta,h}(u)e^{-iut}\Big](-y)\Big\|_{L^{2}}\\
= & \frac{1}{n^{2}\Delta}\frac{\|x\nu\|_{\infty}}{2\pi}\|(m_{\Delta,h}(u)e^{-iut})''\|_{L^{\text{2}}}\|(m_{\Delta,h}(u)e^{-iut})'\|_{L^{2}}\\
\lesssim & \frac{1}{n^{2}\Delta}(1+t^{3})\|(1+|u|)^{\Delta\alpha}\|_{L^{2}(I_h)}^{2}\lesssim\frac{1}{n^{2}\Delta}(1+t^{3})h^{-2\Delta\alpha-1}.
\end{align*}
Moreover, $\xi_{k}$ admits the deterministic bound
\begin{align*}
|\xi_{k}|= & \frac{1}{n\Delta}Y_{k}^{2}\big|\F^{-1}\Big[m_{\Delta,h}(u)e^{-iut}\Big](-Y_{k})\big|\\
\le & \frac{1}{n\Delta}\big|\F^{-1}\big[(m_{\Delta,h}(u)e^{-iut})''\big](-Y_{k})\big|\\
\le & \frac{1}{2\pi n\Delta}\left(\|m_{\Delta,h}''\|_{L^{1}}+2t\|m_{\Delta,h}'\|_{L^1}+t^{\text{2}}\|m_{\Delta,h}\|_{L^{1}}\right)\\
\lesssim & \frac{1}{n\Delta}(1+t^{2})\|(1+|u|)^{\Delta\alpha}\|_{L^{1}(I_h)}\lesssim\frac{1+t^{2}}{n\Delta}h^{-\Delta\alpha-1}.
\end{align*}
Therefore, Bernstein's inequality yields for a constant $c>0$ and
any $\kappa>0$
\[
P\left(|M^\nu_{\Delta,n}(t)|>\kappa\right)\le2\exp\left(-\frac{c\Delta n\kappa^{2}}{(1+t^{3})h^{-2\Delta\alpha-1}+\kappa(1+t^{2})h^{-\Delta\alpha-1}}\right).
\]
Choosing $\kappa=\kappa_{0}(\Delta n)^{-1/2}h^{-\Delta\alpha-1/2}$
for $\kappa_{0}>0$, we conclude
\begin{align*}
P\left(|M^\nu_{\Delta,n}(t)|>\kappa_{0}(\Delta n)^{-1/2}h^{-\Delta\alpha-1/2}\right)\le & 2\exp\left(-\frac{c\kappa_{0}^{2}}{(1+t^{3})(1+\kappa_{0}(n\Delta h)^{-1/2})}\right).\qedhere
\end{align*}
\end{proof}
Proposition~\ref{prop:DensLevyRate} is an immediate consequence of the following result.
\begin{prop}\label{prop:UniformDens}
Let $\alpha,\beta,s,r,R>0,m>4$, let the kernel satisfy (\ref{eq:propKernel})
with order $p\ge s$ and let $U\subset\R$ be a bounded, open set
which is bounded away from zero. Then we have
\begin{enumerate}
\item uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$,
if $\log(n\Delta)/(n\Delta h)\to0$ ,
\[
\sup_{t\in U}|\hat{\nu}_{h}(t)-\nu(t)|=\mathcal{O}_{P,\mathcal{D}^{s}}\Big(h^{s}+\Big(\frac{\log n\Delta}{n\Delta}\Big)^{1/2}h^{-\Delta\alpha-1/2}\Big),
\]
\item uniformly in \textup{$(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)$}
\[
\sup_{t\in U}|\hat{\nu}_{h}(t)-\nu(t)|=\mathcal{O}_{P,\mathcal{E}^{s}}\Big(h^{s}+(n\Delta)^{-1/2}(h^{-1}+\Delta h^{-2}+\Delta^{3/2}h^{-3})e^{r\Delta h^{-\beta}}\Big).
\]
\end{enumerate}
\end{prop}
\begin{proof}
We start with the error decomposition \eqref{eq:DecompDens}. By standard approximation arguments the deterministic error satisfies
$\sup_{t\in U}|B^{\nu}(t)|\lesssim h^{s}$ if $\nu\in C^{s}(U)$ for
an open set $U$ , $s>0$ and if the kernel satisfies (\ref{eq:propKernel})
with order $p\ge s$. Moreover, the assumptions $\supp\F K\subset[-1,1]$
and $x^{p+1}K(x)\in L^{1}(\R)$ imply $\|x^{p+1}K(x)\|_{\infty}<\infty$
which yields
\[
|\sigma^{2}K_{h}(t)|\le\sigma^{2}h^{-1}\sup_{|x|>|t|/h}|K(x)|\lesssim\sigma^{2}|t|^{-s-1}h^{s}.
\]
Since $U$ is bounded away from zero, the previous display gives a
uniform bound on $U$. Using the main stochastic error term $M^\nu_{\Delta,n}$
from (\ref{eq:MainStochErr}), the linearized stochastic error term from \eqref{eq:DecompDens}
can be decomposed similarly to (\ref{eq:DecompLin}) into
\begin{align}
L^\nu_{\Delta,n}(t)=&M^\nu_{\Delta,n}(t)+2\F^{-1}\big[m_{\Delta,h}\psi'(\phi_{\Delta,n}-\phi_{\Delta})'\big](t)\notag\\
&\qquad+{\cal F}^{-1}[m_{\Delta,h}\big(\psi''-\Delta(\psi')^{2}\big)(\phi_{\Delta,n}-\phi_{\Delta})](t).\label{eq:decompLNu}
\end{align}
To derive the appropriate bounds for $R_{\Delta,n}$ and $L^\nu_{\Delta,n}$,
we will distinguish again between the mildly and the severely ill-posed
case.
We start with the severely ill-posed case $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R).$
Using Fubini's theorem and the properties of $m_{\Delta,h}$ as well as $|\psi''(u)|\lesssim1,|\psi'(u)|\lesssim1+|u|,u\in\R,$ and (\ref{eq:phiMom}), we obtain
\begin{align*}
&\E[\sup_{t\in U}|L^\nu_{\Delta,n}(t)|]\\
&\qquad\le \Delta^{-1}\E\Big[\|{\cal F}^{-1}\Big[m_{\Delta,h}(\phi_{\Delta,n}''-\phi_{\Delta}'')\|_{\infty}+2\E\Big[\|\F^{-1}\big[m_{\Delta,h}\psi'(\phi_{\Delta,n}-\phi_{\Delta})'\big]\|_{\infty}\Big]\\
&\qquad\qquad +\E\Big[\|{\cal F}^{-1}[m_{\Delta,h}\big(\psi''-\Delta(\psi'){}^{2}\big)(\phi_{\Delta,n}-\phi_{\Delta})]\|_{\infty}\Big]\\
&\qquad\lesssim \int_{-1/h}^{1/h}\Big(\Delta^{-1}\E[(\phi_{\Delta,n}''(u)-\phi_{\Delta}''(u))^{2}]^{1/2}+\E[(\phi_{\Delta,n}'(u)-\phi_{\Delta}'(u))^{2}]^{1/2}(1+|u|)\\
& \qquad\qquad+\E[(\phi_{\Delta,n}(u)-\phi_{\Delta}(u))^{2}]^{1/2}(1+\Delta(1+|u|)^{2})\Big)\exp(r\Delta|u|^{\beta})\d u\\
&\qquad\le (n\Delta)^{-1/2}\int_{-1/h}^{1/h}\Big(1+\Delta(1+|u|)+\Delta^{1/2}+\Delta^{3/2}(1+|u|)^{2}\Big)\exp(r\Delta|u|^{\beta})\d u\\
&\qquad\lesssim (n\Delta)^{-1/2}(h^{-1}+\Delta h^{-2}+\Delta^{3/2}h^{-3})\exp(r\Delta h^{-\beta}).
\end{align*}
The remainder (\ref{eq:RemainderDens}) is of smaller order
\[
|R_{\Delta,n}|=\mathcal O_{P}\Big(n^{-1}\Delta^{-1/2}\log(h^{-1})^{1+\delta}(h^{-1}+\Delta h^{-2}+\Delta^{3/2}h^{-3})\big)\exp(2r\Delta h^{-\beta})\Big).
\]
Now let us consider $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$,
where we have
\[
|R_{\Delta,n}|=\mathcal{O}_{P}\big((\Delta+\Delta^{3/2}+1)n^{-1}\Delta^{-1/2}\log(h^{-1})^{1+\delta}h^{-2\Delta\alpha-1}\big).
\]
To bound $\sup_{t\in U}|L^\nu_{\Delta,n}(t)|$, we note for the second
and the third term in (\ref{eq:decompLNu}) that similarly to (\ref{eq:nonCritMild})
with the Cauchy--Schwarz inequality and Fubini's theorem
\begin{align*}
& \E\Big[\sup_{t\in U}|L^\nu_{\Delta,n}(t)-M^\nu_{\Delta,n}(t)|\Big]\\
\le & 2\E\Big[\|m_{\Delta,h}\psi'(\phi_{\Delta,n}-\phi_{\Delta})'\|_{L^{1}}\Big]+\E\Big[\|m_{\Delta,h}\big(\psi''-\Delta(\psi'){}^{2}\big)(\phi_{\Delta,n}-\phi_{\Delta})\|_{L^{1}}\Big]\\
\lesssim & 2\|x\nu\|_{L^{2}}\E\big[\|m_{\Delta,h}(\phi_{\Delta,n}-\phi_{\Delta})'\|_{L^{2}}\big]+\|x^{2}\nu\|_{L^{2}}\E\big[\|m_{\Delta,h}(\phi_{\Delta,n}-\phi_{\Delta})\|_{L^{2}}\big]\\
& +\Delta\|x\nu\|_{L^{2}}\E\big[\|m_{\Delta,h}\psi'(\phi_{\Delta,n}-\phi_{\Delta})\|_{L^{2}}\big]\\
\lesssim & n^{-1/2}\Big(2\Delta^{1/2}\|x\nu\|_{L^{2}}+\|x^{2}\nu\|_{L^{2}}+\Delta\|x\nu\|_{L^{2}}\|x\nu\|_{L^{1}}\Big)\|m_{\Delta,h}\|_{L^{2}}\\
\lesssim & n^{-1/2}h^{-\Delta\alpha-1/2}.
\end{align*}
It remains to estimate the main stochastic term $M^\nu_{\Delta,n}$.
Since $U$ is bounded, we find a finite number of points $t_{1},\dots,t_{L_{n}}\in U$
such that $\sup_{t\in U}\min_{l=1,\dots,L_{n}}|t-t_{l}|\le(n\Delta)^{-2}$
and $L_{n}\lesssim(n\Delta)^{2}$. The inequalities by Young and by Cauchy--Schwarz together with $(n\Delta)^{-1}\lesssim1$ yield
\begin{align*}
& \E\Big[\sup_{t\in U}\min_{l=1,\dots,L_{n}}|M^\nu_{\Delta,n}(t)-M^\nu_{\Delta,n}(t_{l})|\Big]\\
\lesssim & (n\Delta)^{-2}\Delta^{-1}\E\Big[\big\|(K_{h}')\ast\big(\F\big[\1_{[-1/h,1/h]}\phi_{\Delta}^{-1}(\phi_{\Delta,n}''-\phi_{\Delta}'')\big]\big)\big\|_{\infty}\Big]\\
\lesssim & (n\Delta)^{-2}\Delta^{-1}h^{-1}\|K'\|_{L^{1}}\int_{-1/h}^{1/h}|\phi_{\Delta}^{-1}|(u)\E[|\phi_{\Delta,n}''-\phi_{\Delta}''|(u)]\d u\\
\lesssim & (n\Delta)^{-2}h^{-\Delta\alpha-2}.
\end{align*}
Together with Markov's inequality and Lemma~\ref{lem:concMain}, defining $T:=\sup_{t\in U}{|t|}$ this
yields for $n\Delta$ sufficiently large
\begin{align*}
& P\Big(\sup_{t\in U}|M^\nu_{\Delta,n}(t)|>\kappa_{0}\Big(\frac{\log(n\Delta)}{n\Delta}\Big)^{1/2}h^{-\Delta\alpha-1/2}\Big)\\
\le & P\Big(\max_{l=1,\dots,L_{n}}|M^\nu_{\Delta,n}(t_{k})|>\frac{\kappa_{0}}{2}\Big(\frac{\log(n\Delta)}{n\Delta}\Big)^{1/2}h^{-\Delta\alpha-1/2}\Big)\\
& +\frac{2}{\kappa_{0}}\big(\frac{n\Delta}{\log(n\Delta)}\big)^{1/2}h^{\Delta\alpha+1/2}\E\Big[\sup_{t\in U}\min_{l=1,\dots,L_{n}}|M^\nu_{\Delta,n}(t)-M^\nu_{\Delta,n}(t_{l})|\Big]\\
\le & 2L_{n}\exp\Big(-\frac{c\kappa_{0}^{2}\log(n\Delta)}{2(1+T^3)(2+\kappa_{0}(\log(n\Delta)/(n\Delta h))^{1/2})}\Big)+o\Big((n\Delta)^{-3/2}(\log n\Delta)^{-1/2}h^{-3/2}\Big)\\
\le & 2\exp\big((2-\tfrac{c}{6}\kappa_{0}^{2}(1+T^3)^{-1})\log(n\Delta)\big)+o(1),
\end{align*}
which converges to zero as $n\Delta\to\infty$ if $\kappa_{0}$ is
chosen sufficiently large.
\end{proof}
\subsubsection{Proof of Theorem~\ref{thm:rateQuantiles}}
In view of the error representation \eqref{eq:errorRep}, the missing ingredient to prove Theorem~\ref{thm:rateQuantiles} is consistency of $\hat q_{\tau,h}^\pm$. We prove it similarly to \cite{dattnerEtAl2013}, but since $\nu$ has no bounded density and we minimize on an unbounded interval the lemma is more involved.
\begin{lemma}
\label{lem:consistency}Let $\alpha,\beta,s,\zeta,r,R>0,m>4,s'\in(-1,0]$
and let $\eta_{n}\downarrow0$ with $\eta_{n}^{-1}\lesssim\log n$.
Suppose the kernel satisfies (\ref{eq:propKernel}) with order $p\ge1$. Then we have
\begin{enumerate}
\item for any bandwidth satisfying $(\log n)^{4}h^{1+s'}\to0$ and $(n\Delta)^{-1}h^{-2\Delta\alpha-1}\to0$
\[
\sup_{(\sigma^{2},\gamma,\nu)\in\mathcal{\tilde{D}}_{\tau}^{s,s'}(\alpha,m,\zeta,\eta_{n},R)}P\big(|\hat{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}|>\delta\big)\to0\quad\text{for all }\delta>0,
\]
\item for any bandwidth satisfying $(\log n)^{4}h^{1+s'}\to0$ and $(\log n)^{4}(n\Delta)^{-1}h^{-2}e^{2r\Delta h^{-\beta}}\to0$
\[
\sup_{(\sigma^{2},\gamma,\nu)\in\mathcal{\tilde{E}}_{\tau}^{s,s'}(\beta,m,\zeta,\eta_{n},r,R)}P\big(|\hat{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}|>\delta\big)\to0\quad\text{for all }\delta>0.
\]
\end{enumerate}
\end{lemma}
\begin{proof}
We adopt the general strategy of the proof of Theorem 5.7 by \citet{vanderVaart1998}
in the classical M-estimation setting. Without loss of generality,
we only consider $\hat{q}_{\tau,h}^{+}$.
\emph{Step 1:} By the H\"older regularity we have $\nu(t)\ge\nu(q_{\tau}^{+})-|\nu(q_{\tau}^{+})-\nu(t)|\ge\frac{1}{R}-R|q_{\tau}^{+}-t|^{1\wedge\alpha}\ge\frac{1}{2R}$
for $|q_{\tau}^{+}-t|\le(2R^{2})^{-(1\vee\alpha^{-1})}$. Without
loss of generality we can assume $\delta\le\delta_0:=(2R^{2})^{-(1\vee\alpha^{-1})}\wedge(\eta_n/2)\wedge\zeta$,
otherwise consider $\delta\wedge\delta_0$.
Using $N(q_{\tau}^{+})=\tau$ and monotonicity of $N$, we obtain
the uniqueness condition
\begin{align}\label{eq:Unique}
\inf_{t>\eta_{n}:|t-q_{\tau}^{+}|\ge\delta}|N(t)-\tau| & \ge\delta\inf_{t>\eta_{n}:|t-q_{\tau}^{+}|\le\delta}\nu(t)\ge\frac{\delta}{2R}.
\end{align}
\emph{Step 2:} We construct an event $A$ with $P(A)\to1$ such that
\begin{equation}\label{eq:ZEst}
\hat N_h(\hat q^+_\tau)-\tau=0\quad\text{on }A.
\end{equation}
Using $N(q^+_\tau)=\tau$, monotonicity of $N$ and \eqref{eq:Unique}, we conclude for $\delta\in(0,\delta_0)$
\[
N(q^+_\tau+\delta)-\tau\le-\frac{\delta }{2R}<0<\frac{\delta }{2R}\le N(q^+_\tau-\delta)-\tau.
\]
Proposition~\ref{prop:DistFunct} implies that
\begin{equation}\label{eq:A5}
A:=\big\{\forall \eta\in\{q_\tau-\delta,q_\tau+\delta\}:|\hat N_h(\eta)-N(\eta)|\le\frac{\delta }{4R}\big\}\quad\text{satisfies}\quad P(A)\to1.
\end{equation}
We conclude on $A$ that $\hat N_h(q_\tau+\delta)<\tau< \hat N_h(q_\tau-\delta)$. Since $\hat N_h$ admits a derivative, namely $\hat\nu_h$, it is continuous and thus there is an intermediate point $\xi_h\in(q^+_\tau-\delta,q^+_\tau+\delta)$ such that $\hat N_h(\xi_h)=\tau$. Since $\hat{q}_{\tau,h}^{+}$ minimizes $|\hat{N}_{h}(\bull)-\tau|$
on the interval $(\eta_{n},\infty)$ and $q_{\tau}-\delta\in(\eta_{n},\infty)$ for $\eta_n$ and $\delta$ sufficiently small, we obtain \eqref{eq:ZEst}.
\emph{Step 3:} We infer from Steps~1 and 2
\begin{align}
P\big(|\hat{q}_{\tau,h}^{+}-q_{\tau}^{+}|>\delta\big) & \le P\big(|N(\hat{q}_{\tau,h}^{+})-\tau|\ge\delta/(2R)\big)\nonumber \\
& =P\big(|N(\hat{q}_{\tau,h}^{+})-\hat{N}_{h}(\hat{q}_{\tau,h}^{+})|\ge\delta/(2R)\big)+o(1)\nonumber \\
& \le P\big(\sup_{t\in(\eta_{n},\infty)}|N(t)-\hat{N}_{h}(t)|\ge\delta/(2R)\big)+o(1).\label{eq:ConsistencyBoundLevy}
\end{align}
Hence, it remains to show uniform consistency of $\hat{N}_{h}(t)$.
Applying the error decomposition~(\ref{eq:Decomp}) $|N(t)-\hat{N}_{h}(t)|\le|B_{n}(t)|+|S_{n}(t)|+|V_{n}(t)|$
and the estimates in Lemma~\ref{lem:vola} and Proposition~\ref{prop:bias}
we obtain
\begin{align*}
\sup_{t\in(\eta_{n},\infty)}|V_{n}(t)|\lesssim & \eta_{n}^{-4}h,\qquad\sup_{t\in(\eta_{n},\infty)}|B_{n}(t)|\lesssim\eta_{n}^{-4}h^{1+s'}.
\end{align*}
According to Lemma~\ref{lem:Remainder} and Proposition~\ref{prop:variance},
the stochastic error term can be decomposed into $S_{n}(t)=M_{\Delta,n}(t)+(L_{\Delta,n}-M_{\Delta,n})(t)+R_{n}(t),$
where $\sup_{|t|\ge\eta_{n}}|L_{\Delta,n}-M_{\Delta,n}|(t)+|R_{n}|(t)$
is of the order claimed in Proposition~\ref{prop:variance}. For
$(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,m,U,r,R)$ the main
stochastic error term is uniformly bounded as well. For the case $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,m,U,R)$
it remains to apply Proposition~\ref{prop:mainStochErr} on an appropriate
grid. For $\kappa\in(0,1)$ we define a grid $v_{l}=-\kappa^{-1}+l\kappa$
with $l=0,\dots,L:=2\lfloor\kappa^{-2}\rfloor$. Set $t_l:=\sign (v_l)(|v_l|\vee\eta_n)$ for $l=0,\dots,L$. Then
\begin{align*}
\sup_{|t|\in(\eta_{n},\infty)}|M_{\Delta,n}(t)|\le & \sup_{l=0,\dots,L}|M_{\Delta,n}(t_{l})|+\sup_{t\in(\eta_{n},\infty)}\min_{l=0,\dots,L}|M_{\Delta,n}(t)-M_{\Delta,n}(t_{l})|.
\end{align*}
Noting that $\|g_{t}-g_{s}\|_{L^{1}}\le|t-s|/(s^{2}\wedge t^{2})$
and $\|g_{t}-g_{s}\|_{L^{1}}\le2\int_{|t|\wedge|s|}^{\infty}x^{-2}\d x\sim|t|^{-1}\vee|s|^{-1}$,
increments of $M_{\Delta,n}(t)$ can be estimated using Plancherel's
identity and Fubini's theorem
\begin{align*}
& \E\Big[\sup_{t\in(\eta_{n},\infty)}\min_{l=0,\dots,L}|M_{\Delta,n}(t)-M_{\Delta,n}(t_{l})|\Big]\\
\le & \frac{1}{2\pi\Delta}\E\Big[\sup_{t\in(\eta_{n},\infty)}\min_{l=0,\dots,L}\Big|\int\F[g_{t}-g_{t_{l}}](-u)m_{\Delta,h}(u)(\phi_{\Delta,n}''(u)-\phi_{\Delta}''(u))\d u\Big|\Big]\\
\le & \frac{1}{2\pi\Delta}\sup_{t\in(\eta_{n},\infty)}\min_{l=0,\dots,L}\|g_{t}-g_{t_{l}}\|_{L^{1}}\int_{-1/h}^{1/h}|m_{\Delta,h}(u)|\E[(\phi_{\Delta,n}''(u)-\phi_{\Delta}''(u))^{2}]^{1/2}\d u\\
\lesssim & (n\Delta)^{-1/2}\sup_{t\in(\eta_{n},\infty)}\min_{l=0,\dots,L}\|g_{t}-g_{t_{l}}\|_{L^{1}}\|(1+|u|)^{\Delta\alpha}\|_{L^{1}(I_h)}\\
\lesssim & (n\Delta)^{-1/2}\eta_{n}^{-2}\kappa h^{-\Delta\alpha-1}.
\end{align*}
Choosing $\kappa=\eta_{n}^{2}(n\Delta)^{-1/2}$, Markov's inequality
and Proposition~\ref{prop:mainStochErr} yield for $\Delta n$ sufficiently
large and some constant $c>0$
\begin{align*}
& P\Big(\sup_{|t|\in(\eta_{n},\infty)}|M_{\Delta,n}(t)|>\delta\Big)\\
\le & P\Big(\sup_{l=0,\dots,L}|M_{\Delta,n}(t_{j})|>\frac{\delta}{2}\Big)+\frac{2}{\delta}\E\Big[\sup_{t\in(\eta_{n},\infty)}\min_{l=0,\dots,L}|M_{\Delta,n}(t)-M_{\Delta,n}(t_{l})|\Big]\\
\le & 2(L+1)\exp\Big(-c\delta^{2}\eta_{n}^{3}n\Delta\|(1+|u|)^{\Delta\alpha-1}\|_{L^{2}(I_h)}^{-2}\Big)+(n\Delta)^{-1/2}\eta_{n}^{-2}\kappa h^{-\Delta\alpha-1}\\
\le & 2\eta_{n}^{-4}n\Delta\exp(-c\delta^{2}\eta_{n}^{3}n\Delta h^{2\Delta\alpha})+(n\Delta)^{-1}h^{-\Delta\alpha-1}\to0,
\end{align*}
owing to $n\Delta h^{2\Delta\alpha+1}\to\infty$.
\end{proof}
$\,$
\begin{proof}[Proof of Theorem \ref{thm:rateQuantiles}]
Proposition~\ref{prop:DistFunct} shows that the numerator in the
error representation (\ref{eq:errorRep}) is of the claimed order.
Moreover, it holds for any $\delta>0$
\[
P(|\hat{\nu}_{h}(\xi^{\pm})-\nu(q_{\tau}^{\pm})|>\delta)\le P(\sup_{|t|<\zeta}|\hat{\nu}_{h}(q_{\tau}^{\pm}+t)-\nu(q_{\tau}^{\pm})|>\delta)+P(|\hat{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}|\ge\zeta),
\]
where the first term converges to zero by Proposition~\ref{prop:UniformDens}
and the second one tends to zero by Lemma~\ref{lem:consistency}.
Therefore, the denominator in (\ref{eq:errorRep}) can be written
as
\[
\hat{\nu}_{h}(\xi^{\pm})=\nu(q_{\tau}^{\pm})+o_{P}(1).\qedhere
\]
\end{proof}
\subsection{Proofs for Section \ref{sec:FinancialQuantile}}
To prove Lemma \ref{lem:GaussProc}, we
apply some entropy arguments. To fix the notation, we define
for any (pseudo-)metric $d$ on $\R$ the covering number $N(r,A,d)$
as the smallest number of $d$-balls with radius $r>0$ which is necessary
to cover a subset $A\subset\R$. For $v>0$ the entropy integral is
defined as
\[
J(v,A,d):=\int_{0}^{v}\sqrt{\log N(r,A,d)}\d r,
\]
which is finite for any $v$ if $N(r,A,d)$ grows polynomially in
$r^{-1}$.
\begin{proof}[Proof of Lemma~\ref{lem:GaussProc}]
It\^o's isometry yields
\begin{align}
\Var(\Phi_{n}(u)) =\E[|\Phi_{n}(u)|^{2}]&=n^{-1}u^{2}(u^{2}+1)\E\Big[\Big|\int e^{iux-x}\rho(x)\d W(x)\Big|^{2}\Big]\notag\\
&=n^{-1}u^{2}(u^{2}+1)\|e^{-x}\rho(x)\|_{L^{2}}^{2}\label{eq:VarGaussProc}
\end{align}
and similarly for $\Phi_{n}^{(1)}$ and $\Phi_{n}^{(2)}$ as defined
in (\ref{eq:GaussProc}). From It\^o's isometry and dominated convergence
we conclude that $\Phi_{n}^{(k)},k=1,2,$ are the first and second order
$L^{2}(P)$-derivatives of $\Phi_{n}$. The intrinsic covariance metric
of the Gaussian process $(\int e^{iux-x}x^{k}\rho(x)\d W(x))_{u}$
is given by
\[
d^{(k)}(u,v):=\E\Big[\Big|\int(e^{iux}-e^{ivx})x^{k}e^{-x}\rho(x)\d W(x)\Big|^{2}\Big]^{1/2}.
\]
Using $\int|x|^{m}e^{-2x}\rho^{2}(x)\d x<\infty$, the entropy integrals
$J(\infty,[-U,U],d^{(k)})$ can be bounded exactly as in the proof
of Proposition~1 by \citet{soehl2014} and are of order $\sqrt{\log U}.$
Dudley's theorem \citep[e.g.][Prop. 3.18]{massart2007} yields then
$\E[\|\Phi_{n}^{(k)}\|_{L^{\infty}[-U,U]}]\lesssim n^{-1/2}U^{2}\sqrt{\log U})$.
\end{proof}
\subsubsection{Convergence rates}
Since the proof strategy is the same for the observation schemes
in Sections~\ref{sec:quantileEst} and \ref{sec:FinancialQuantile},
we will concentrate here on the differences. The estimation error can be decomposed
into $\tilde{N}_{t}(t)-N(t)=B_{n}(t)+\tilde{S}_{n}(t)+V_{n}(t)$ where
$B_{n}(t)$ and $V_{n}(t)$ are given in (\ref{eq:Decomp}) and only
the stochastic error
\[
\tilde{S}_{n}(t)=\int g_{t}(x)\F^{-1}\Big[\Big(\psi''(u)-\tilde{\psi}_{n}''(u)\Big)\F K(hu)\Big](x)\d x
\]
has a different probabilistic structure. We can apply Lemma~\ref{lem:vola}
and Proposition~\ref{prop:bias} to estimate $B_{n}(t)$ and $V_{n}(t)$. For the sake of brevity, we will frequently write
\[
(\phi_{T}^{-1}(\tilde{\phi}_{T,n}-\phi_{T}))''
=\phi_T^{-1}\Phi_n^{(2)}+2(\phi_T^{-1})'\Phi_n^{(1)}+(\phi_T^{-1})''\Phi_n.
\]
This equality is justified in $L^2(P)$-sense, but should merely be understood as notational convention. Linearizing the stochastic error term, we define analogously to (\ref{eq:DecompLin})
\begin{align*}
\tilde{L}_{T,n}(t):= & -\frac{1}{T}\int g_{t}(x){\cal F}^{-1}[\F K(h\bull)(\phi_{T}^{-1}(\tilde{\phi}_{T,n}-\phi_{T}))''](x)\d x.
\end{align*}
Since the $\Phi_{n}^{(k)}$ are almost surely bounded and $\F K$ has
compact support, $\tilde{L}_{T,n}$ is almost surely well defined. The remainder $\tilde S_n-\tilde L_{T,n}$ will be bounded on the event
\begin{equation}
\Omega_{n,h}:=\Big\{\inf_{u\in I_h}|\tilde{\phi}_{T,n}(u)|\ge n^{-1/2}h^{-2}(\log h^{-1})\Big\}.\label{eq:Omega}
\end{equation}
The order of the remainder in the next lemma corresponds exactly to Lemma~\ref{lem:Remainder} taking the bound from Lemma~\ref{lem:GaussProc} into account.
\begin{lemma}\label{lem:RemainderOptions}
If $\int|x|^{m}e^{-2x}\rho^{2}(x)\d x<\infty$
for some $m>4$, then for any sequence $h=h_{n}$ satisfying
$n^{-1/2}h^{-2}(\log h^{-1})\|\phi_{T}^{-1}\|_{L^{\infty}(I_{h})}\to0$
as $n\to\infty$ it holds $P(\Omega_{n,h})\to 1$ and uniformly for all L\'evy triplets $(\sigma^2,\gamma,\nu)$
\begin{align*}
&\sup_{u\in I_h}\left|\tilde{\psi}_{n}''(u)-\psi''(u)-T^{-1}(\phi_{T}^{-1}(\tilde{\phi}_{T,n}-\phi_{T}))''(u)\right|\\
& =\mathcal O_P\Big((1+\|\psi'\big\|_{L^{\infty}(I_h)}^{2})n^{-1}h^{-4}\log(h^{-1})\|\phi_{T}^{-1}\|_{L^{\infty}(I_{h})}^{2}\Big).
\end{align*}
\end{lemma}
\begin{proof}
$P(\Omega_{n,h})\to 1$ follows immediately from Theorem~1 by \cite{kappusReiss2010}, cf. Lemma 5.1 by \cite{dattnerEtAl2013}. To bound $\tilde{\phi}_{T,n}^{-1}-\phi_{T}^{-1}$, we apply the argument by \citet{neumann1997} and Lemma~\ref{lem:GaussProc}. We obtain on $\Omega_{n,h}$
\begin{align}
\|\tilde{\phi}_{T,n}^{-1}-\phi_{T}^{-1}\|_{L^\infty(I_h)}\1_{\Omega_{n,h}}
&\le\sup_{u\in I_h}\Big(\frac{|\Phi_{n}(u)|}{|\phi_{T}(u)|^{2}}+n^{1/2}h^2|\log h|^{-1}\frac{|\Phi_{n}(u)|^{2}}{|\phi_{T}(u)|^{2}}\Big)\label{eq:NeumannTrick}\\
&=\mathcal O_P\big(n^{-1/2}h^{-2}|\log h|^{1/2}\|\phi_{T}^{-1}\|^2_{L^{\infty}(I_{h})}\big).\notag
\end{align}
A straight forward computation shows on $\Omega_{n,h}$
\begin{align}
&T\big(\tilde{\psi}_{n}''-\psi''\big)
=\frac{\tilde\phi_{T,n}''}{\tilde \phi_{T,n}}-\frac{\phi_T''}{\phi_T}-\Big(\frac{\tilde\phi_{T,n}'}{\tilde\phi_{T,n}}\Big)^2+\Big(\frac{\phi_T'}{\phi_T}\Big)^2\notag\\
=&\frac{\tilde\phi_{T,n}''-\phi_T''}{\phi_T}
-\Big(\frac{\tilde\phi_{T,n}'}{\tilde\phi_{T,n}}+\frac{\phi_T'}{\phi_T}\Big)\frac{\tilde\phi_{T,n}'-\phi_T'}{\phi_T}
+\Big(\frac{\tilde\phi_{T,n}''}{\tilde\phi_{T,n}}-\Big(\frac{\tilde\phi_{T,n}'}{\tilde\phi_{T,n}}+\frac{\phi_T'}{\phi_T}\Big)\frac{\tilde\phi_{T,n}'}{\tilde\phi_{T,n}}\Big)\frac{\phi_T-\tilde\phi_{T,n}}{\phi_T}\notag\\
=&\frac{\tilde\phi_{T,n}''-\phi_T''}{\phi_T}
-2\Big(\frac{\phi_{T}'}{\phi_{T}}\Big)\frac{\tilde\phi_{T,n}'-\phi_T'}{\phi_T}
+\Big(\frac{\phi_{T}''}{\phi_{T}}-2\frac{\phi_T'}{\phi_T}\frac{\phi_{T}'}{\phi_{T}}\Big)\frac{\phi_T-\tilde\phi_{T,n}}{\phi_T}+R_n\notag\\
=&\phi_T^{-1}\Phi_n^{(2)}+2(\phi_T^{-1})'\Phi_n^{(1)}+(\phi_T^{-1})''\Phi_n+R_n\label{eq:remainderUgly}
\end{align}
where $R_n$ is the sum of all second order terms. Using \eqref{eq:NeumannTrick}, formulas \eqref{eq:PhiDerivatives} as well as Lemma~\ref{lem:GaussProc}, we obtain the claimed order of $R_n$.
\end{proof}
Let us study the linearized stochastic error term. To apply the Lepski method later we need a sharp bound on the variance of $\tilde L_{T,n}$. With the auxiliary functions, $u\in\R$,
\begin{align*}
\chi_t^{(0)}(u)&:=\F g_{t}(-u)\F K(hu)\big(u(u-i)(\phi_{T}^{-1})''(u)+2(2u-i)(\phi_T^{-1})'(u)+2\phi_T^{-1}(u)\big),\\
\chi_t^{(1)}(u)&:=\F g_{t}(-u)\F K(hu)\big(2u(iu+1)(\phi_{T}^{-1})'(u)+(4iu+2)\phi_{T}^{-1}(u)\big),\\
\chi_t^{(2)}(u)&:=u(i-u)\F g_{t}(-u)\F K(hu)\phi_{T}^{-1}(u)
\end{align*}
we define
\begin{align}\label{eq:DefSigma}
\Sigma_{n,h}(t):=&\frac{1}{2\pi n^{1/2}T}\Big(\|x^{2}e^{-x}\rho(x)\|_{\infty}\|\chi_t^{(2)}\|_{L^2}\notag\\
&\qquad\quad\qquad+\|xe^{-x}\rho(x)\|_{\infty}\|\chi_t^{(1)}\|_{L^2}+\|e^{-x}\rho(x)\|_{\infty}\|\chi_t^{(0)}\|_{L^2}\Big).
\end{align}
\begin{lemma}\label{lem:LinFin}
If $\int(1+|x|)^{4}e^{-2x}\rho^{2}(x)\d x<\infty$,
then $\tilde{L}_{T,h}(t)$ is centered normal. Supposing additionally
$\|(1\vee x^{2})e^{-x}\rho\|_{\infty}\lesssim1$, it holds
\begin{align}
\E[|\tilde{L}_{T,n}(t)|^{2}]^{1/2}
\le \Sigma_{n,h}(t)
\lesssim n^{-1/2}(t^{-1}\vee t^{-2})\big\|(1+ |u|)|\phi_{T}(u)|^{-1}(1+\psi'(u)^{2})\big\|_{L^{2}(I_{h})}.\nonumber
\end{align}
\end{lemma}
\begin{proof}
Since the $\Phi_{n}^{(k)}$ are almost surely bounded, we can apply Plancherel's
identity which yields
\begin{align*}
\E[|\tilde{L}_{T,n}(t)|^{2}] & =\frac{1}{(2\pi T)^2}\E\Big[\Big|\int\F g_{t}(-u)\F K(hu)\phi_{T}^{-1}(u)\Phi_{n}^{(2)}(u)\d u\\
& \qquad\qquad\quad+2\int\F g_{t}(-u)\F K(hu)(\phi_{T}^{-1})'(u)\Phi_{n}^{(1)}(u)\d u\\
& \qquad\qquad\quad+\int\F g_{t}(-u)\F K(hu)(\phi_{T}^{-1})''(u)\Phi_{n}(u)\d u\Big|^{2}\Big].
\end{align*}
By continuity and boundedness of $\Phi_{n}^{(k)},k=0,1,2,$ the integral
in $u$ can be approximated with Riemann sums. We conclude first that
$\tilde{L}_{T,h}(t)$ is normally distributed, cf. Section 6.2 in
\citet{soehl2014}, and second that we can exchange the deterministic
integral and the stochastic integral due to the construction of the
Wiener integral as $L^{2}(P)$-limit. Together with It\^o's isometry
and Plancherel's identity we obtain
\begin{align*}
&\E[|\tilde{L}_{T,n}(t)|^{2}]^{1/2}\\
= & \frac{1}{n^{1/2}T}\E\Big[\Big|\int\Big(x^{2}\F^{-1}\chi_t^{(2)}(-x)+x{\cal F}^{-1}\chi_t^{(1)}(-x)+\F^{-1}\chi_t^{(0)}(-x)\Big)e^{-x}\rho(x)\d W(x)\Big|^{2}\Big]^{1/2}\\
= & \frac{1}{n^{1/2}T}\Big(\int\Big|x^{2}\F^{-1}\chi_t^{(2)}(-x)+x{\cal F}^{-1}\chi_t^{(1)}(-x)+\F^{-1}\chi_t^{(0)}(-x)\Big|^{2}e^{-2x}\rho^{2}(x)\d x\Big)^{1/2}\\
\le & \frac{1}{2\pi n^{1/2}T}\Big(\|x^{2}e^{-x}\rho(x)\|_{\infty}\|\chi_t^{(2)}\|_{L^2}+\|xe^{-x}\rho(x)\|_{\infty}\|\chi_t^{(1)}\|_{L^2}+\|e^{-x}\rho(x)\|_{\infty}\|\chi_t^{(0)}\|_{L^2}\Big).
\end{align*}
The formulas (\ref{eq:PhiDerivatives}) and $|\F g_{t}(u)|\lesssim(t^{-1}\vee t^{-2})(1+|u|)^{-1}$ yield the claimed asymptotic bound.
\end{proof}
Now, we can conclude convergence rates for the distribution function
estimator $\tilde{N}_{h}$.
\begin{prop}
Suppose $\|(1\vee x^{2})e^{-x}\rho\|_{\infty}\lesssim1$ and $\int(1+|x|)^{m}e^{-2x}\rho^{2}(x)\d x<\infty$
for some $m>4$. Let $U\subset\R$ be an open set and $\alpha,\beta,s,r,R>0.$
Let the kernel satisfy \eqref{eq:propKernel} with order $p\ge s+1$.
\begin{enumerate}
\item If $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,2,U,R)$, then
$|\tilde{N}_{h}(t)-N(t)|=\mathcal{O}_{P}\big(n^{-(s+1)/(2s+2\Delta\alpha+5)}\big)$
for $h=h_{n}=n^{-1/(2s+2\Delta\alpha+5)}$.
\item If $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,2,U,r,R)$, then
$|\tilde{N}_{h}(t)-N(t)|=\mathcal{O}_{P}\big((\log n)^{-(s+1)/\beta}\big)$
for $h=h_{n}=(\frac{1}{2r}\log n)^{-1/\beta}$.
\end{enumerate}
\end{prop}
\begin{proof}
As in the proof of Proposition \ref{prop:DistFunct} we have
\begin{align}
|\tilde{N}_{h}(t)-N(t)|\le & |B_{n}(t)|+|\tilde{S}_{n}(t)|+|V_{n}(t)|\lesssim h^{s+1}+|\tilde{S}_{n}(t)|.\label{eq:ErrDistFunctFin}
\end{align}
Lemmas~\ref{lem:RemainderOptions} and \ref{lem:LinFin} yield for
the stochastic error term
\begin{align*}
|\tilde{S}_{n}(t)|& \le|\tilde{L}_{n}(t)|+|\tilde{S}_{n}(t)-\tilde{L}_{n}(t)|\\
& =\mathcal O_P\Big( n^{-1/2}(1+\|\psi'\big\|_{L^{\infty}(I_h)}^{2})\\
&\qquad\qquad\times\Big(\big\|(1+|u|)\phi_{T}(u)|^{-1}\big\|_{L^{2}(I_h)}+n^{-1/2}h^{-4}(\log h^{-1})\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}^{2}\Big)\Big).
\end{align*}
In situation (i) we use integrability of $x\nu$ to obtain
\begin{align}
|\tilde{N}_{h}-N|(t)=\mathcal{O}_{P}\Big( & h^{s+1}+n^{-1/2}\big(\|(1+|u|)^{\Delta\alpha+1}\|_{L^{2}(I_h)}+n^{-1/2}(\log h^{-1})h^{-2\Delta\alpha-4}\big)\Big)\notag\\
=\mathcal{O}_{P}\Big( & h^{s+1}+n^{-1/2}\big(h^{-\Delta\alpha-3/2}+n^{-1/2}(\log h^{-1})h^{-2\Delta\alpha-4}\big)\Big).\label{eq:ErrNmild}
\end{align}
Therefore, the optimal bandwidth $h_{n}=n^{-1/(2s+2\Delta\alpha+5)}$
yields the claimed rate.
For exponentially decaying characteristic functions in case (ii) we
infer by $\|\psi'\big\|_{L^{\infty}(I_h)}\lesssim h^{-1}$
\begin{align}
|\tilde{N}_{h}-N|(t)= & \mathcal{O}_{P}\Big(h^{s+1}+n^{-1/2}\big(h^{-2}\|(|u|+1)e^{r|u|^{\beta}}\|_{L^{2}(I_h)}+n^{-1/2}h^{-6}(\log h^{-1})e^{rh^{-\beta}}\big)\Big)\notag\\
= & \mathcal{O}_{P}\Big(h^{s+1}+n^{-1/2}h^{-7/2}(1+n^{-1/2}h^{-5/2}(\log h^{-1}))e^{rh^{-\beta}}\Big),\label{eq:ErrNsev}
\end{align}
leading to the rate optimal choice $h_{n}=(\frac{1}{2r}\log n)^{-1/\beta}$.
\end{proof}
Proposition $\ref{prop:UniformDensRates}$ on the density estimator
$\tilde{\nu}_{h}$ is an immediate consequence from the following
proposition.
\begin{prop}\label{prop:UniformDensFin}
Suppose $\|(1\vee x^{2})e^{-x}\rho\|_{\infty}\lesssim1$ and $\int(1+|x|)^{m}e^{-2x}\rho^{2}(x)\d x<\infty$
for some $m>4$. Let the kernel satisfy (\ref{eq:propKernel})
with order $p\ge s$, let $U\subset\R$ be a bounded, open set which
is bounded away from zero and let $\alpha,\beta,r,R>0$.
Then we have
\begin{enumerate}
\item for any $h\downarrow0$ satisfying $n^{-1/2}h^{-\Delta\alpha-5/2-\delta}\to0$
for some $\delta>0$ we have uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{D}^{s}(\alpha,2,U,R)$
\begin{align*}
\sup_{t\in U}t^{2}|\tilde{\nu}_{h}(t)-\nu(t)| & =\mathcal O_{P,\mathcal{D}^{s}}\Big( h^{s}+n^{-1/2}(\log n)^{1/2}h^{-\Delta\alpha-5/2}\Big),
\end{align*}
\item for any $h\downarrow0$ satisfying $n^{-1/2}e^{(r+\delta)h^{-\beta}}\to0$
for some $\delta>0$ we have uniformly in $(\sigma^{2},\gamma,\nu)\in\mathcal{E}^{s}(\beta,2,U,r,R)$
\[
\sup_{t\in U}t^2|\tilde{\nu}_{h}(t)-\nu(t)|=\mathcal O_{P,\mathcal{E}^{s}}\Big( h^{s}+n^{-1/2}(\log n)^{1/2}h^{-9/2}e^{rh^{-\beta}}\Big).
\]
\end{enumerate}
\end{prop}
\begin{proof}
As in the proof of Proposition~\ref{prop:UniformDens} we deduce
from Lemma~\ref{lem:RemainderOptions} that
\begin{align*}
\sup_{t\in U}t^2|\tilde{\nu}_{h}(t)-\nu(t)|=&\mathcal O(h^{s})+\sup_{t\in U}t^2|\tilde{L}_{T,n,\nu}(t)|\\
&\qquad+\mathcal O_P\Big((1+\|\psi'\big\|_{L^{\infty}(I_h)}^{2})n^{-1}h^{-5}\log(h^{-1})\|\phi_{\Delta}^{-1}\|_{L^{\infty}(I_h)}^{2}\Big)
\end{align*}
with the linearized stochastic error term
\begin{align*}
\tilde{L}_{T,n,\nu}(t):=-\frac{1}{Tt^2}\F^{-1}\Big[\F K(h\bull)\Big(\frac{\tilde{\phi}_{T,n}-\phi_{T}}{\phi_{T}}\Big)''\Big](t).\\
\end{align*}
We estimate
\begin{align*}
\E\big[\sup_{t\in U}|t^2\tilde{L}_{T,n,\nu}(t)|\Big] & \le T^{-1}\E\Big[\sup_{t\in U}|{\cal F}^{-1}[m_{T,h}\Phi_n^{(2)}](t)|]+\E[\sup_{t\in U}|{\cal F}^{-1}[m_{T,h}\psi'\Phi_n^{(1)}](t)|]\\
& \qquad+\E[\sup_{t\in U}|{\cal F}^{-1}[m_{T,h}(\psi''-T(\psi')^2)\Phi_n](t)|\Big]\\
& =:E_{1}+E_{2}+E_{3}.
\end{align*}
Since all three terms can be estimated analogously, we limit ourselves on
$E_{3}$ which has the largest variance. We estimate uniformly in $t\in U$ with use of Plancherel's
identity, Fubini's theorem and It\^o's isometry
\begin{align}
&\Var\big(\F^{-1}[m_{T,h}(\psi''-T(\psi')^2)\Phi_n](t)\big) \notag\\
=&\frac{1}{2\pi}\E\Big[\Big|\int e^{-itu}m_{T,h}(u)(\psi''(u)-T\psi'(u)^2)\Phi_n(u)\d u\Big|^{2}\Big]\nonumber \\
=&n^{-1}\E\Big[\Big|\int\F^{-1}[m_{T,h}(u)(\psi''(u)-T\psi'(u)^2)u(u-i)e^{-itu}](-x)\rho(x)\d W(x)\Big|^{2}\Big]\nonumber \\
=&n^{-1}\int\big|\F^{-1}[m_{T,h}(u)(\psi''(u)-T\psi'(u)^2)u(u-i)e^{-itu}](-x)\rho(x)\big|^{2}\d x\nonumber \\
\lesssim& n^{-1}\int(u^{2}+u^{4})|m_{T,h}(u)|^{2}(1+|\psi'(u)|^2)^2\d u=:v(n,h).\label{eq:varDenfin}
\end{align}
Analogously, we can estimate the distance in the intrinsic norm for
any $\delta\in(0,1/2)$
\begin{align*}
d(s,t)^{2}:= & \E\Big[\Big|{\cal F}^{-1}[m_{T,h}(\psi''-T(\psi')^2)\Phi_n](t)-{\cal F}^{-1}[m_{T,h}(\psi''-T(\psi')^2)\Phi_n](s)\Big|^{2}\Big]\\
= & n^{-1}\int(1+u^{4})|m_{T,h}(u)|^{2}(1+|\psi'(u)|^2)^2|e^{-itu}-e^{-isu}|^2\d u\\
\lesssim & |t-s|^{2\delta}n^{-1}\int(1+u^{4+2\delta})|m_{T,h}(u)|^{2}(1+|\psi'(u)|^2)^2\d u\\
=: & |t-s|^{2\delta}c_\delta(n,h)
\end{align*}
and thus the covering number is of the order $N(r,U,d)\lesssim(c_\delta(n,h)/r)^{1/(2\delta)}.$
Consequently, the entropy integral can be bounded by
\begin{align}
J\Big(\sqrt{v(n,h)},U,d\Big) & \lesssim\int_{0}^{\sqrt{v(n,h)}}\big(\log c_\delta(n,h)+\log r^{-1}\big)^{1/2}\d r\nonumber \\
& \lesssim v(n,h)^{1/2}\big(\log c_\delta(n,h)+\log v(n,h)^{-1}\big)^{1/2}.\label{eq:entropyIntDens}
\end{align}
Using this entropy bound, Dudley's theorem \citep[e.g.][Prop. 3.18]{massart2007}
yields
\[
\E\Big[\sup_{t\in U}|\F^{-1}[m_{T,h}(\psi''-T(\psi')^2)\Phi_n](t)|\Big]\lesssim v(n,h)^{1/2}\big(\log c_\delta(n,h)+\log v(n,h)^{-1}\big)^{1/2}.
\]
Now we can plug in the different assumptions on the decay of $\phi_{T}$
(in particular $\log c_\delta(n,h)$ is smaller than $0$ for $\delta$ sufficiently
small).
\end{proof}
To finally prove Theorem~\ref{thm:quantileFin}, we can argue as
for Theorem \ref{thm:rateQuantiles}. The only ingredient which remains be shown is uniform convergence $\sup_{|t|>\eta}|\tilde{N}_{h}(t)-N(t)|=o_P(1)$
for the rate optimal bandwidth $h=h_{n}$. Since the remainder $|\tilde{S}_{n}(t)-\tilde{L}_{n}(t)|$
can be bounded uniformly in $t$ using Lemma~\ref{lem:RemainderOptions},
it suffices to show:
\begin{lemma}\label{lem:uniformLtilde}
If $\|(1\vee x^{2})e^{-x}\rho\|_{\infty}\lesssim1$ and $\int(1+|x|)^{m}e^{-2x}\rho^{2}(x)\d x<\infty$
for some $m>4$, then it holds uniformly for all $(\sigma^2,\gamma,\nu)$
\[
\E[\sup_{|t|\ge\eta}|\tilde{L}_{T,n}(t)|]\lesssim\eta^{-2}n^{-1/2}(\log n)^{1/2}\Big(\int|m_{T,h}(u)|^{2}\big(1+|\psi'(u)|^{2}\big)(1+u^{4})\d u\Big)^{1/2}.
\]
\end{lemma}
\begin{proof}
Let us estimate the covering number of $\R\setminus(-\eta,\eta)$
with respect to the intrinsic metric of the process $\tilde{L}_{T,n}(t)$.
Similarly to Proposition~\ref{prop:UniformDensFin} we infer
\begin{align*}
d(s,t): & =\E[|\tilde{L}_{T,n}(t)-\tilde{L}_{T,n}(s)|^{2}]^{1/2}\\
& \lesssim n^{-1/2}\Big(\int|\F[g_{t}-g_{s}](-u)m_{T,h}(u)|^{2}\big(1+|\psi'(u)|^{2}\big)(1+u^{4})\d u\Big)^{1/2}\\
& \lesssim\|g_{t}-g_{s}\|_{L^{1}}n^{-1/2}\Big(\int|m_{T,h}(u)|^{2}\big(1+|\psi'(u)|^{2}\big)(1+u^{4})\d u\Big)^{1/2}\\
& =:\|g_{t}-g_{s}\|_{L^{1}}c(n,h).
\end{align*}
As in Lemma \ref{lem:consistency}, we see that the covering numbers
are polynomial:
\[
N\big(r,\R\setminus(-\eta,\eta),d\big)\lesssim\eta_{n}^{-4}c(n,h)^{-2}r^{-2}.
\]
With the variance bound $\Sigma_{n,h}:=\sup_{|t|>\eta}\Sigma_{n,h}(t)$ from \eqref{eq:DefSigma} Dudley's theorem and Lemma~\ref{lem:LinFin} yield
\begin{align*}
\E[\sup_{|t|\ge\eta}|\tilde{L}_{T,n}(t)|]
&\lesssim J(\Sigma_{n,h},\R\setminus(-\eta,\eta),d)\\
&\lesssim\Sigma_{n,h}\big(\log(\eta_{n}^{-2}c(n^{-1/2},h))+\log \Sigma_{n,h}^{-1}\big)^{1/2}.\qedhere
\end{align*}
\end{proof}
\subsubsection{Adaptive method}
With the previous results at hand the proof of Theorem~\ref{thm:quantileFinAdapt} is quite similar to the one of Theorem~3.2 by \cite{dattnerEtAl2013} and thus we omit the details. In particular, the bandwidth set $\mathcal{B}_{n}$ fulfills analogous properties as their construction, cf. their Lemma~3.1. Due to $P(\Omega_{n,h})\to1$ for the minimal bandwidth and with $\Omega_{n,h}$ from (\ref{eq:Omega}), it suffices to bound all terms on the complement $\Omega_{n,h}^{c}$. Using (\ref{eq:errorRep}), \eqref{eq:ZEst} and \eqref{eq:ErrDistFunctFin}, the estimation error of $\tilde{q}_{\tau,h}$ can be bounded by
\begin{align}
|\tilde{q}_{\tau,h}^{\pm}-q_{\tau}^{\pm}|\le & \frac{|B_{n,h}(q_{\tau}^{\pm})|+|\tilde{S}_{n,h}(q_{\tau}^{\pm})|+|V_{n,h}(q_{\tau}^{\pm})|}{|\tilde{\nu}_{n,h}(\xi^{\pm})|}\le\frac{Dh^{s+1}+|\tilde{S}_{n,h}(q_{\tau}^{\pm})|}{|\tilde{\nu}_{n,h}(\xi^{\pm})|}\label{eq:errorDecompTilde}
\end{align}
with probability converging to one and with a deterministic constant $D>0$ involving the bias and the error due to $\sigma^2$. We can verify with use of Proposition~\ref{prop:UniformDensFin}, (\ref{eq:ConsistencyBoundLevy}) and Lemmas~\ref{lem:RemainderOptions}
and \ref{lem:uniformLtilde} for $\eta\in(0,1)$,
\begin{equation}\label{eq:denominator}
P\Big(\max_{h\in\mathcal{B}_{n}}\sup_{\xi^{\pm}\in[q_{\tau}^{\pm}\wedge\tilde{q}_{\tau,h}^{\pm},q_{\tau}^{\pm}\vee\tilde{q}_{\tau,h}^{\pm}]}|\tilde{\nu}_{n,h}(\xi^{\pm})-\nu(q_{\tau}^{\pm})|>\eta\nu(q_{\tau}^{\pm})\Big)\to0.
\end{equation}
To conclude that $\tilde{V}_{n}^{\pm}(h)$ from \eqref{eq:V} is an appropriate upper bound for $|\tilde{S}_{n,h}(t)|/|\tilde{\nu}_{n,h}(\xi^{\pm})|$
in (\ref{eq:errorDecompTilde}), we have to control $\tilde{S}_{n,h}(q_{\tau}^{\pm})$. We again decompose it into linearization $\tilde{L}_{n,h}(q_{\tau}^{\pm})$ and
remainder. Since $\tilde{L}_{n,h}(q_{\tau}^{\pm})$ is centered and normally distributed
with variance bounded by $\Sigma_{n,h}^{2}(q_\tau^\pm)$ from (\ref{eq:DefSigma}),
the Gaussian concentration and $|\mathcal{B}_{n}|\lesssim\log n$
yield for any $\delta>0$
\[
P\Big(\exists h\in\mathcal{B}_{n}:|\tilde{L}_{n,h}(q_{\tau}^{\pm})|>(1+\delta)\sqrt{2\log\log n}\Sigma_{n,h}(q_\tau^\pm)\Big)\to0.
\]
For the remainder, one can show, using (\ref{eq:remainderUgly}) pointwise, $\E[|\tilde{S}_{n,h}(q_\tau^\pm)-\tilde{L}_{n,t}(q_\tau^\pm)|\1_{\Omega_{n,h}}]\lesssim\Sigma_{n,h}(q_\tau^\pm)^{2}h^{-1}$
and thus we conclude for any $\delta>0$
\[
P\Big(\exists h\in\mathcal{B}_{n}:|\tilde{S}_{n,h}(q_\tau^\pm)|>(1+\delta)\sqrt{2\log\log n}\Sigma_{n,h}(q_\tau^\pm)\Big)\to0,
\]
provided that $(\log n)\Sigma_{n,h}(q_\tau^\pm)h^{-1}\to0$. Noting
that the order of $\Sigma_{n,h}(q_\tau^\pm)h^{-1}$ is the same as the order
of the stochastic error of $\tilde{\nu}_{n,h}$, this condition is
satisfied for all $h\in\mathcal{B}_{n}$ by construction. In the next
step, we show that $\Sigma_{n,h}(q_\tau^\pm)$ is reasonably estimated by $\tilde{\Sigma}_{n,h}^\pm$
from (\ref{eq:SigmaTilde}).
\begin{lemma}
In the situation of Theorem~\ref{thm:quantileFinAdapt} we have for any sequence
$h\downarrow0$ satisfying $\inf_{|u|\le 1/h}|\phi_{T}(h)|>n^{-1/2}h^{-2}\log h^{-1}$
that
\[
|\tilde{\Sigma}_{n,h}^\pm-\Sigma_{n,h}(q_\tau^\pm)|=\mathcal O_P\big((\log h^{-1})^{-1}\Sigma_{n,h}(q_\tau^\pm)\big).
\]
\end{lemma}
\begin{proof}
Without loss of generality we only consider $q_\tau^+$. The triangle inequality yields
\begin{align*}
|\tilde{\Sigma}_{n,h}^+-\Sigma_{n,h}(q_\tau^+)|
\le & \frac{1}{2\pi n^{1/2}T}\Big(\|x^{2}e^{-x}\rho(x)\|_{\infty}\|\tilde\chi_{\tilde q_{\tau,h}^+}^{(2)}-\chi_{q_\tau^+}^{(2)}\|_{L^2}\\
&\quad+\|xe^{-x}\rho(x)\|_{\infty}\|\tilde\chi_{\tilde q_{\tau,h}^+}^{(1)}-\chi_{q_\tau^+}^{(1)}\|_{L^2}+\|e^{-x}\rho(x)\|_{\infty}\|\tilde\chi_{\tilde q_{\tau,h}^+}^{(0)}-\chi_{q_\tau^+}^{(0)}\|_{L^2}\Big).\\
\end{align*}
Hence, it suffices to show
\begin{equation}\label{eq:EstSig}
\|\tilde\chi_{\tilde q_{\tau,h}^+}^{(k)}-\chi_{q_\tau^+}^{(k)}\|_{L^2}=\mathcal O_P\big((\log h^{-1})^{-1}\|\chi_{q_\tau^+}^{(k)}\|_{L^2}\big),\quad\text{for }k=0,1,2.
\end{equation}
Let us start with $k=2$ where we have
\begin{align*}
&\int\big|\tilde\chi_{\tilde q_{\tau,h}^+}^{(2)}(u)-\chi_{q_\tau^+}^{(2)}(u)\big|^2\d u\\
=&\int(u^2+u^4)|\F K(hu)|^2\big|\F g_{\tilde q_{\tau,h}^+}(-u)\tilde\phi_{T,n}^{-1}(u)-\F g_{q_\tau^+}(-u)\phi_{T}^{-1}(u)\big|^2\d u\\
\le&\int(u^2+u^4)|\F K(hu)|^2|\phi_T(u)|^{-2}\big|\F g_{\tilde q_{\tau,h}^+}(-u)-\F g_{q_{\tau}^+}(-u)\big|^2\d u\\
&\quad+\int(u^2+u^4)|\F K(hu)|^2|\F g_{\tilde q_{\tau,h}^+}(-u)|^2\big|\tilde\phi_{T,n}^{-1}(u)-\phi_{T}^{-1}(u)\big|^2\d u\\
=:&T_1+T_2.
\end{align*}
Using $(g_t-g_s)(x)=x^2\1_{(s,t]}$ for $0<s<t$ and $|\F g_t(u)|\sim(1+|u|)^{-1}$, the first integral can be estimated by
\begin{align*}
T_1&\le\|g_{\tilde q_{\tau,h}^+}-g_{q_{\tau}^+}\|_{L^1}^2\int(u^2+u^4)|\F K(hu)|^2|\phi_T(u)|^{-2}\d u\\
&\le\eta_n^{-4}|\tilde q_{\tau,h}^+-q_{\tau}^+|^2h^{-2}\|\chi_{q_\tau^+}^{(2)}\|_{L^2}^2.
\end{align*}
Applying \eqref{eq:ErrNmild}, \eqref{eq:ErrNsev} and \eqref{eq:denominator}, we conclude $T_1=\mathcal O_P((\log h^{-1})^{-2}\|\chi_{q_\tau^+}^{(2)}\|_{L^2}^2)$.
To bound $T_2$, we estimate with \eqref{eq:NeumannTrick} on $\Omega_{n,h}$
\begin{align*}
T_2\lesssim&\eta_n^{-4}\int(u^2+1)|\F K(hu)|^2|\big|\tilde\phi_{T,n}^{-1}(u)-\phi_{T}^{-1}(u)\big|^2\d u\\
\le&\eta_n^{-4}\int(u^2+1)|\F K(hu)|^2|\phi_{T}(u)|^{-4}\big(|\Phi_{n}(u)|^{2}+n(u^{2}+u^{4})^{-1}|\Phi_{n}(u)|^{4}\big)\d u
\end{align*}
Using the estimate (\ref{eq:VarGaussProc}), $\Phi_{n}$ is a centered
normal random variable with variance smaller than $n^{-1}u^{2}(u^{2}+1)\|e^{-x}\rho\|_{L^{2}}^{2}$.
Therefore,
\begin{align*}
T_{2}& =\mathcal O_P\Big(n^{-1}\eta_n^{-4}\int u^2(u^{2}+1)^2|\F K(hu)|^2|\phi_{T}(u)|^{-4}\big)\d u\Big)\\
& =\mathcal O_P\Big( n^{-1}\eta_n^{-4}h^{-4}\sup_{|v|\le1/h}|\phi_{T}(v)|^{-2}\|\chi_{q_\tau^+}^{(2)}\|_{L^2}^2\Big)\\
& =\mathcal O_P\Big((\log h^{-1})^{-2}\|\chi_{q_\tau^+}^{(2)}\|_{L^2}^2\Big).
\end{align*}
We conclude \eqref{eq:EstSig} for $k=2$. For $k=0,1$ similar calculations
apply using
\begin{align*}
(\tilde{\phi}_{T,n}^{-1}-\phi_{T}^{-1})' & =\frac{-\Phi_{n}^{(1)}+T\psi'\Phi_{n}}{\tilde{\phi}_{T,n}^{2}}+T\psi'(\phi_{T}^{-1}-\tilde{\phi}_{T,n}^{-1}),\\
(\tilde{\phi}_{T,n}^{-1}-\phi_{T}^{-1})'' & =\frac{-\Phi_{n}^{(2)}+T\psi''\Phi_{n}+T\psi'\Phi_{n}^{(1)}}{\tilde{\phi}_{T,n}^{2}}\\
&\qquad-\frac{2}{\tilde{\phi}_{T,n}^{2}}\big(1-\frac{\Phi_{n}}{\tilde{\phi}_{T,n}}\big)\big(T\psi'+\frac{\Phi_{n}^{(1)}}{\phi_{T}}\big)\big(-\Phi_{n}^{(1)}+T\psi'\Phi_{n}\big)\\
& \qquad+T\big(\psi''(\phi_{T}^{-1}-\tilde{\phi}_{T,n}^{-1})+\psi'(\phi_{T}^{-1}-\tilde{\phi}_{T,n}^{-1})'\big).\qedhere
\end{align*}
\end{proof}
With these preparations at hand, we can proceed exactly as in \cite[Sect. 5.2]{dattnerEtAl2013} to see that $\tilde{h}^{\pm}$
mimics the oracle bandwidth which balances the deterministic error
$Dh^{s+1}$ and the stochastic error $\tilde{S}_{n,h}$ in the numerator
of (\ref{eq:errorDecompTilde}). Details are omitted.
\bibliographystyle{apalike}
|
\section{Introduction}
The string theory is a very valuable tool for us to understand properties of the black hole.
One of the great successes in this field is the statistical explanation of the Bekenstein-Hawking entropy for the certain extremal and the near extremal black hole.
Different types of black holes have been studied well and all of them support the explanation for the Bekenstein-Hawking entropy since the first success in \cite{17}.
Recent works in this field have focused on the question that in stationary spacetime allowing the event horizon is not Killing horizon \cite{1}.
Some former works \cite{4,5,6} may forbid this construction.
However, from a physical perspective, there are two different surface gravities at the horizons where they connect two asymptotic spaces, which indicate different temperatures.
In thermodynamic theory, this is a description about a steady heat stream between two infinite heat reservoirs that preserve a fixed temperature gradient when the time is under evolution \cite{1}.
In article \cite{1}, a flowing horizon was constructed by a very impressive way. Basically, they build up the construction with a Schwarzschlid metric that is a static black hole solution.
However, in this article we extend the former solution to more general conditions.
We consider a Kerr metric to construct the flow, and it is easy to extend to a Kerr-Newman metric condition.
Even though the descriptions of horizon flows motivated by AdS/CFT \cite{2,18,19,20,21,22,Huang:2010yg,Huang:2,Fang-Fang:2013rra} have been studied well, as the same reason in \cite{1}, in our construction we do not need a negative cosmological constant to hold the system steadily.
We begin our construction from such a picture: a very thin black string falls perpendicularly and smoothly into a very large black hole.
Their radius are denoted by $r_{bs}$ and $r_{hs}$ respectively.
Both the black string and the black hole rotate with the same angular velocity $\alpha$ and the axes of the rotation is the direction of the falling black string flow.
Since we have $r_{bh}\gg r_{bs}$, given that the surface gravity is inverse ratio to the radius, we know that the surface gravity of black string is much larger than the surface gravity of black hole.
When the string is falling into the black hole horizon without any external interference, we can anticipate that the two horizons fuse smoothly into each other.
By this construction, the black hole can accumulate mass from the falling string and, as the consequence, the black hole must grow in size irreversibly.
However, when we take the limit $r_{bh}\to \infty$ and hold $r_{bs}$ fixed, and then just pay attention on events that happen in the intersected region we can get rid of the effect of the growing of the black hole.
The black hole horizon then becomes a Rindler-like infinite horizon because of the rotation.
If the angular velocity is zero, the horizon is exactly the Rindler horizon.
By changing to the rest frame of the falling black string, the acceleration horizon vanishes. Then we have a configuration of a stationary black string.
Alternatively, if we take a stationary black string and study it from a frame that accelerates along the direction of the string, what we observe is that a string free falls into an acceleration horizon.
We will construct the event horizon for such acceleration observers and study how the rotation will affect the result, and then make a comparison with the non-rotation result.
The arrangement of this paper are: in Section 2 we will construct a Kerr black string flow and study the properties of the construction.
In Section 3 we will study the black string flow and its equilibrium condition, and then extend to a more general model. The last section is our conclusion and outlook.
\section{Construction of Kerr black string flow}
Configurations of higher dimensional rotating and charged black holes obtained in string theory have been studied a lot on a micro level \cite{23,24,25}.
Unlike these constructions above, we follow \cite{1} to build up a new metric.
\subsection{Horizon of black string flow}
In article \cite{1} the authors give their metric as follows
\begin{equation}
ds^2=-f\left(r\right)dt^2+dz^2+\frac{dr^2}{f\left(r\right)}+r^2 d\Omega_{n+1},
\end{equation}
it is a black string flow without rotation in $D=n$ spacetime dimensions.
Generally we can generalize metric (2.1) to a new expression with rotation
\begin{eqnarray}
ds^2 &=& -\left(1-\frac{2Mr}{\rho ^2}\right)dt^2+\frac{\rho ^2}{\Delta}dr^2+\rho ^2 d\theta ^2 +[\left(r^2+\alpha ^2\right)sin^2 \theta +\frac{2Mr\alpha ^2 sin^4 \theta}{\rho ^2}]d\phi ^2 \nonumber\\
&-&\frac{4Mr\alpha sin^2 \theta}{\rho ^2}dtd\phi + Fdz^2,
\end{eqnarray}
where
\begin{eqnarray}
\rho ^2 = & r^2 + \alpha ^2 cos^2 \theta, \\
\Delta = & r^2 - 2Mr + \alpha ^2,
\end{eqnarray}
and F is the metric function that should have a positive sign outside the black string horizon and be finite at $r \to \infty$ when $\alpha$ is fixed.
The metric (2.2) is in the rest frame of the free-falling and rotating black string in D=5 dimensions.
When the black string is missing $\left( M=0 \right)$, we have the null surfaces $\left(t=z+t_0\right)$ that are the acceleration horizons with different constant $t_0$.
One condition we need to guarantee is that when it is far from the string and at the same time the rotation can be neglected, the null rays can change to the conventional Rindler horizon, i.e.,
\begin{equation}
\frac{\dot{t}}{\dot{z}}\to 1 ~ ~ ~ and ~ ~ ~ \dot{r} \to 0 ~ ~ ~ for ~ ~ ~ r \to \infty,
\end{equation}
where the dot indicates the derivative of an affine parameter $\lambda$.
By transforming the metric into Boyer-Lindquist coordinate, we have
\begin{equation}
ds^2=-\frac{\Delta}{\rho ^2}dT^2 +\frac{\rho ^2}{\Delta}dR^2 +\rho ^2 d\Omega +FdZ^2,
\end{equation}
with
\begin{equation}
d\Omega =d \Theta ^2 + sin^2 \Theta d\Phi.
\end{equation}
The relations between new and old metric are
\begin{eqnarray}
dT &=& dt-\alpha sin^2 \theta d\phi, \\
dR &=& dr, \\
d\Phi &=& \frac{\left(r^2 +\alpha ^2\right)d\phi-\alpha dt}{\rho ^2},\\
d\Theta &=& d\theta, \\
dZ &=& dz .
\end{eqnarray}
Since the affine parameter $\lambda$ is also valid in new coordinate system, we have the null path
\begin{equation}
-\frac{\Delta}{\rho ^2}\dot{T}^2 +\frac{\rho ^2}{\Delta}\dot{R}^2+F \dot{Z}^2 =0,
\end{equation}
and
\begin{equation}
\dot{T}=\frac{\epsilon \rho ^2}{\Delta}, ~ ~ ~ \dot{Z}=\frac{p}{F}.
\end{equation}
In equation (2.14), $\epsilon$ and $p$ are two integration constants because of the isometries generated by $\partial _T$ and $\partial _Z$. The details are given in appendix A.
From equations (2.5), (2.8) and (2.12) we know the relations between $\epsilon$ and $p$
\begin{eqnarray}
\frac{dT}{d\lambda}&=&\frac{dt}{d\lambda}-\alpha sin^2\theta \frac{d\phi}{d\lambda},\\
\frac{dZ}{d\lambda}&=&\frac{dz}{d\lambda},
\end{eqnarray}
and because
\begin{equation}
\frac{dt}{d \lambda}=\frac{dz}{d\lambda}~ ~ ~ when~ ~ ~ r\to \infty,
\end{equation}
we get
\begin{equation}
\epsilon + \alpha sin^2\theta \frac{d\phi}{d\lambda}=p.
\end{equation}
We can neglect the contributions of F and $\frac{\rho^2}{\Delta}$ because when $r \to \infty$, both of them approximate to 1.
\subsection{Null hypersurface and the analysis}
Now from the equations (2.14) and (2.18), we have the representation of $\dot{Z}$
\begin{equation}
\dot{Z}=\frac{\Delta \dot{T}+\rho^2 A_\theta}{F\rho^2},
\end{equation}
with
\begin{equation}
A_\theta=\alpha sin^2 \theta \frac{d\phi}{d\lambda}.
\end{equation}
Using equations (2.13) and (2.19), we have
\begin{equation}
\dot{R}=\pm\sqrt{\left(\frac{\Delta}{\rho^2}\right)^2\dot{T}^2-\frac{\Delta}{F\rho^2}\left(A_\theta+\frac{\Delta}{\rho^2}\dot{T}\right)^2}.
\end{equation}
Because both of the $\dot{Z}$ and $\dot{R}$ can be represented by $\dot{T}$, we can deduce an equation that consists of $\dot{Z}$, $\dot{T}$ and $\dot{R}$,
\begin{equation}
\dot{T}=\dot{Z}\mp\frac{\Delta \dot{T} -F\rho^2\dot{T} +\rho^2 A_\theta}{F \sqrt{\Delta^2 \dot{T}^2 -\frac{\Delta\left(\Delta \dot{T} +\rho^2 A_\theta\right)^2}{F\rho ^2 }}}\dot{R}.
\end{equation}
Equation (2.22) can be proved by putting equations (2.19) and (2.21) into equation (2.22).
Then the null hypersurfaces ruled by the one-form equation are
\begin{equation}
dT=dZ\mp\frac{\Delta dT -F\rho^2dT +\rho^2 \hat{A}_\theta}{F \sqrt{\Delta^2 dT^2 -\frac{\Delta\left(\Delta dT +\rho^2\hat{A}_\theta\right)^2}{F\rho ^2 }}}dR,
\end{equation}
where $\hat{A}_\theta=\alpha sin^2 \theta d\phi.$
Back to the old coordinate system we get
\begin{equation}
dt-\alpha sin^2 \theta d\phi =dz\mp\frac{\left(dt-\alpha sin^2\theta d\phi \right)\Delta -F\left(dt-\alpha sin^2 \theta d\phi\right)\rho^2 +\rho^2 \hat{A}_\theta }{F\sqrt{\left(dt-\alpha sin^2\theta d\phi\right)^2 \Delta^2-\frac{\Delta\left(\left(dt-\alpha sin^2 \theta d\phi\right)\Delta+\rho^2 \hat{A}_\theta\right)^2 }{F\rho^2 } } }dr.
\end{equation}
When $\alpha =0$, Kerr metric will return to Schwarzschlid metric. By setting F=1 and $\frac{\Delta}{\rho^2}=f$ for convenience, the null hypersurfaces become
\begin{equation}
dt=dz\pm\frac{\sqrt{1-f}}{f}dr.
\end{equation}
These are the null hypersurfaces ruled by geodesics in \cite{1}.
Because of the missing symmetry, equations (2.23) and (2.24) both have extremely complicated resolutions.
For convenience and no losing generality, we first set $F=1$ and $M=1$, then just study the event horizon at $\theta=0$.
In this way we get the new one-form equation
\begin{equation}
dT=dZ\mp\frac{\left(\Delta dT -\rho^2 dT\right) }{\sqrt{\Delta^2 dT^2-\Delta^3 \rho^{-2}dT^2 } }dR.
\end{equation}
Benefiting from $\theta =0$ the old coordinate system is equal to new one and as a consequence we get
\begin{equation}
dt=dz\mp\frac{\left(r^2+\alpha^2\right)^{\frac{1}{2}}\sqrt{2r}}{r^2-2r+\alpha^2 }dr.
\end{equation}
Then we have
\begin{equation}
t=z+t_0\mp\int \frac{\left(r^2+\alpha^2\right)^{\frac{1}{2}}\sqrt{2r}}{r^2-2r+\alpha^2 }dr,
\end{equation}
each value of $t_0$ results in a corresponding null hypersurface that ends at different value in the null coordinate in future null infinity.
Clearly any of them can be regarded as our event horizon.
Here, for convenience, we denote the null hypersurface with $t_0=0$ as the horizon.
This is not a stationary horizon: the action of $\partial_t$ can change $t_0$, therefore it does not map a hypersurface onto itself but onto another one instead.
The explicit form of the integral result of $r$ in (2.28) is given it in the appendix B.
The surfaces $H_f$ are plotted in figure 1.
\begin{figure}
\centering
\includegraphics[width=2in ,clip]{event-horizon.eps}
\includegraphics[width=2in ,clip]{event-horizon3.eps}
\hfill
\caption{\label{fig:1}Null hypersurfaces $H_f$ of the black string flow are plotted from equation (2.28) in (t, r, z) space for $F=M=\alpha =1$ and $\theta=0$. The left one is the out-going hypersurface which means we choose the positive sign in equation (2.28). The right one is the in-going hypersurface. The red curves are the constant-t section.}
\end{figure}
In these pictures, $M=\alpha$ indicates that they are extreme cases, the inner event horizon and outer event horizon meet at $r=1$.
A constant imaginary part occurs when $r<1$, we use the magnitude of this complex number to get these figures.
\subsection{Null geodesic congruence}
Without losing generality, we just consider the negative term in equation (2.23) which corresponds to the outgoing hypersurface.
Benefiting from the freedom to scale $\lambda$, we can set $\epsilon =p-A_\theta=1$.
Then from the equation (2.14), we have
\begin{equation}
\dot{T}=\frac{\rho ^2}{\Delta}, ~ ~ ~ \dot{Z}=\frac{1+A_\theta}{F}.
\end{equation}
According to equation (2.22), we have
\begin{equation}
\frac{\rho^2}{\Delta}=\frac{1+A_\theta}{F}-\frac{\rho^2 -F\frac{\rho^4}{\Delta} +\rho^2 A_\theta}{F\sqrt{\Delta}\rho \sqrt{\frac{\rho^2 }{\Delta}-\frac{\left(1 + A_\theta\right)^2}{F }}}\dot{R}.
\end{equation}
This gives us
\begin{equation}
\dot{R}=\sqrt{\frac{-\Delta+F\rho^2-2\Delta A_\theta-\Delta A^2 _\theta}{F\rho^2}}.
\end{equation}
Because of $dR=dr$, equation (2.31) can be rewritten in $\left( t, r, z \right)$ coordinate system easily
\begin{equation}
\frac{dr}{d\lambda}=\sqrt{\frac{C_2r^2+C_1r+C_0}{F\left(r^2+\alpha^2cos^2\theta\right)} },
\end{equation}
with
\begin{eqnarray}
C_0 &=&\alpha^2\left(cos^2\theta F-\left(A_\theta+1\right)^2\right),\nonumber\\
C_1 &=& 2M\left(1+A_\theta\right)^2,\nonumber \\
C_2 &=&F-1-2A_\theta-A^2 _\theta.\nonumber
\end{eqnarray}
Again we consider the situation in figure 1.
By setting $\theta=0$ and $ F=M=\alpha=1$, we get
\begin{equation}
\frac{dr}{d\lambda}=\sqrt{\frac{2r}{r^2+1} },
\end{equation}
The function $\lambda\left(r\right)$ is plotted in figure 2.
More specific calculations see appendix C.
\begin{figure}
\centering
\includegraphics[width=2in ,clip]{geodesic.eps}
\includegraphics[width=2in ,clip]{geodesic2.eps}
\hfill
\caption{\label{fig:2}The behavior of function $\lambda \left(r\right)$. The picture on the left depicts the function when r is large. And on the right we show the behavior of the function when r is close to 0, and one should notice that 0 is a singularity. When $r \to 0 ,$ we get $\lambda \to \infty$. And as $\lambda$ grows, r moves toward to $\infty$.}
\end{figure}
\subsection{Event horizon and ergosurface}
In Kerr black hole, the region between the outer event horizon and the Killing horizons is known as the ergosphere.
Inside the ergosphere, any objects have to rotate with the black hole, but at the same time they are free to move toward or away from the event horizon.
When we consider the Kerr black string flow, the same thing happens on hypersurface $H_f$.
According to equation (2.33) and figure 2, we find that all null rays on $H_f$ must go towards $r \to \infty$ as the $\lambda$ grows.
This means that any timelike trajectory that remains within the bounded values of r will cross $H_f$ at last.
As a consequence, any observers that remain within a finite range of the black string will fall across $H_f$ eventually.
As the same as the story in Kerr black hole, any observers inside this surface cannot remain static but are dragged along with the string, so this is a particular ergoshpere that the Kerr string flow can distinguish from an ordinary Kerr black hole.
In our construction the black string is accelerating, and then the ergoregion grows.
For any observes who wants to avoid to fall across this $H_f$, an acceleration in the z direction is not enough, they also need to move out towards $r \to \infty$.
Even though the hypersurface $H_f$ we consider here is a special case, it is straightforward to extend this analysis to any $\theta$.
\section{Out-of-equilibrium flow}
Let us consider the vector
\begin{eqnarray}
\frac{d}{d\lambda} &=& \dot{T}\frac{\partial}{\partial T}+\dot{Z}\frac{\partial}{\partial Z}+\dot{R}\frac{\partial}{\partial R}\\
&=& \frac{\rho ^2}{\Delta}\frac{\partial}{\partial T}+\frac{\left(1+A_\theta\right)}{F}\frac{\partial}{\partial Z}+\sqrt{\frac{-\Delta+F\rho^2-2\Delta A_\theta-\Delta A^2 _\theta}{F\rho^2}}\frac{\partial}{\partial R}.\nonumber
\end{eqnarray}
This is an affine generator of the null geodesic congruence.
It is convenient to consider the following non-affine generator\footnote{According to the definition in \cite{carroll}, any parameter $\lambda$, which relates to the proper time $\tau$ in this way: $\lambda=a\tau+b$ where $a$ and $b$ are constants, is an affine parameter. It is obvious that $\frac{1}{2}\frac{\Delta}{\rho^2}$ is not a constant, so parameter $l$ is not an affine parameter and the generator of $l$ is not an affine generator, too.} of the future horizon
\begin{eqnarray}
l &=& \frac{1}{2} \frac{\Delta}{\rho^2}\frac{d}{d\lambda} \\
&=& \frac{1}{2}\left(\frac{\partial}{\partial T}+\frac{\Delta\left(1+A_\theta\right)}{\rho^2F}\frac{\partial}{\partial Z}+\frac{\Delta}{\rho^2}\sqrt{\frac{-\Delta+F\rho^2-2\Delta A_\theta-\Delta A^2 _\theta}{F\rho^2}}\frac{\partial}{\partial R}\right).\nonumber
\end{eqnarray}
The reason we normalize the generator into equation (3.2) is these generators can recover to the black string horizon and the acceleration horizon near and far from the black string respectively.
The surface gravity $\kappa_{\left(l\right)}$ of $l$ is defined in this form
\begin{equation}
\nabla_l l=\kappa_{\left(l\right)}l.
\end{equation}
Given that $\lambda$ is an affine parameter, we have
\begin{eqnarray}
\kappa_{\left(l\right)}&=&\frac{1}{2}\frac{d}{d\lambda}\left[\frac{\Delta}{\rho^2}\right]\nonumber\\
&=&\frac{-M\alpha^2 cos^2 \theta-\alpha^2r+r\alpha^2cos^2\theta+Mr^2}{\left(\alpha^2cos^2\theta+r^2\right)^2}\frac{dr}{d\lambda}.
\end{eqnarray}
The specific calculations are showed in Appendix D.
According to equation (2.32), we have
\begin{equation}
\kappa_{\left(l\right)}=\left(\frac{-M\alpha^2 cos^2 \theta-\alpha^2r+r\alpha^2cos^2\theta+Mr^2}{\left(\alpha^2cos^2\theta+r^2\right)^2}\right)\sqrt{\frac{C_2r^2+C_1r+C_0}{F\left(r^2+\alpha^2cos^2\theta\right)} }.
\end{equation}
This surface gravity diminishes monotonically to 0 when $r$ is large enough.
And when $\alpha =0$ the Kerr black string flow returns to Schwarzschlid black string flow. By setting $F=1$ we get the surface gravity
\begin{equation}
\kappa_{\left(l\right)}=\frac{1}{2r}\left(\frac{2M}{r}\right)^{\frac{3}{2}},
\end{equation}
which coincides with the result in \cite{1}.
There is a chance that by setting a specific $F$ we can make the surface gravity $\kappa=0$.
Namely, putting concrete coefficients $C_i~\left(i=0,1,2\right)$ into equation (3.5), we have
\begin{equation}
F=\frac{\left(-2Mr+r^2+\alpha^2\right)\left(1+2A_\theta+A^2 _\theta\right)}{r^2+\alpha^2 cos^2\theta}.
\end{equation}
Now we set $M=1$ and $\frac{d\phi}{d\lambda}=\alpha$ to concretely give different expressions of equation (3.7).
We plot this function in figure 3 with four different conditions ($\alpha=$1, 0.75, 0.5, 0,25 respectively).
\begin{figure}
\centering
\includegraphics[width=1.6in ,clip]{Fa10.eps}
\includegraphics[width=1.6in ,clip]{Fa075.eps}
\\
\includegraphics[width=1.6in ,clip]{Fa05.eps}
\includegraphics[width=1.6in ,clip]{Fa025.eps}
\hfill
\caption{\label{fig:3} In these pictures we show the function $F\left(r,\alpha,\theta\right)$ with four different $\alpha$. The first one is the extreme condition that $M=\alpha=1$, the event horizon locates at $r=1$. Other three pictures are ordinary conditions the red lines denote their outer horizons $r_H$, when $r>r_H$, the function $F$ is positive.}
\end{figure}
Our previous analysis about metric (2.6) can be extended to a more general form, for example, the charged and rotating black string.
The metric can be written in this form
\begin{equation}
ds^2=-\mathcal{T}dT^2+\mathcal{Z}dZ^2+\frac{dR^2}{\mathcal{R}}+R^2d\Omega ,
\end{equation}
where $\mathcal{T},\mathcal{Z}$ and $ \mathcal{R}$ are general functions of angular momentum J, charge Q and the space coordinates $\left( r, \theta, \phi \right)$. The metric (3.8) is represented in Boyer-Lindquist coordinate by applying equations (2.7)-(2.12).
Here we restrict ourselves to $\mathcal{T}<\mathcal{Z}$.\footnote{When $\mathcal{T}=\mathcal{Z}$ the string worldsheet is Lorentz-invariant. A more circumstantial analysis sees \cite{1}.}
Similar to the deduction and calculation of equation (2.23), we achieve the null hypersurface
\begin{equation}
dT=dZ+\frac{\mathcal{Z}dT-\mathcal{T}dT-\hat{A}_\theta}{\mathcal{Z}\sqrt{\mathcal{R}}\sqrt{\mathcal{T}dT^2-\frac{\left(\mathcal{T}dT+\hat{A}_\theta\right)^2}{\mathcal{Z}}}}dR,
\end{equation}
where $\hat{A}_\theta = \alpha sin^2 \theta d\phi$.
Similar to the deductions of equations (2.29) and (2.31), the congruences in terms of the affine parameter $\lambda$ are
\begin{eqnarray}
dZ&=&\left(1+\hat{A}_\theta\right)\mathcal{Z}^{-1}d\lambda+d\xi,\\
dR&=&\sqrt{\mathcal{R}}\left(\frac{1}{\mathcal{T}}-\frac{\left(1+\hat{A}_\theta\right)^2}{\mathcal{Z}}\right)^{1/2}d\lambda ,
\end{eqnarray}
here the letter $\xi$ is to label the different null rays.
Then, following the beginning of section 3, we can give the non-affine null geodesic generator
\begin{eqnarray}
l &=&\frac{1}{2}\mathcal{T}^{-1}\frac{d}{d\lambda}\\
&=& \frac{1}{2}\left(\frac{\partial}{\partial T}+\frac{\mathcal{T}\left(1+A_\theta\right)}{\mathcal{Z}}\frac{\partial}{\partial Z}+\sqrt{\mathcal{R}}\sqrt{\frac{1}{\mathcal{T}}-\frac{\left(1+A_\theta\right)^2}{\mathcal{Z}}}\frac{\partial}{\partial R}\right)\nonumber,
\end{eqnarray}
where $A_\theta=\alpha sin^2 \frac{d\phi}{d\lambda}$.
And similar to appendix D, because $\lambda$ is an affine parameter, we have the surface gravity $\kappa_{\left(l\right)}$
\begin{eqnarray}
\kappa_{\left(l\right)}&=&\frac{1}{2}\frac{d}{d\lambda}\left[\mathcal{T}\right]\nonumber\\
&=&\frac{1}{2}\frac{dR}{d\lambda}\frac{\partial \mathcal{T}}{\partial R}\\
&=&\frac{1}{2}\sqrt{\mathcal{R}}\sqrt{\frac{1}{\mathcal{T}}-\frac{\left(1+A_\theta\right)^2}{\mathcal{Z}}}\frac{\partial \mathcal{T}}{\partial R}.\nonumber
\end{eqnarray}
When the rotation decreases to 0, we get the ordinary result
\begin{equation}
\kappa_{\left(l\right)}=\frac{1}{2}\sqrt{\left(\mathcal{Z}-\mathcal{T}\right)\frac{\mathcal{R}}{\mathcal{T}\mathcal{Z}}}\frac{\partial \mathcal{T}}{\partial r}.
\end{equation}
Again this is the result in the static spacetime.
In equation (3.13), we can get an equilibrium condition
\begin{equation}
\mathcal{Z}=\left(1+A_\theta\right)^2\mathcal{T}.
\end{equation}
This extreme limit will happen when we consider the black strings with basic string charge.
The charged string flow describes a string with a specific excitation down to a large black hole.
Even if the string charge can decrease the temperature of the black string, only at equation (3.15) can it be in thermal equilibrium.
Above the extreme situation, the string excitation is much hotter than that of the black hole, and the system seems to have a different behavior from the configuration which the string excitation is in thermal equilibrium with a finite temperature horizon (a worldsheet approach \cite{a,b,c,d} or a blackfold approach \cite{e,f}).
In this formation, when the black string does not attain to extreme, it will dump energy to the black hole.
\section{Conclusion and outlook}
In this paper, we generally give and analyze a model of the rotating black string flow in dimension D=5, and extend this solution to a charged and rotating black sting flow, and study their equilibrium conditions. In these constructions, the smooth intersection between the black hole and black string is in the axisymmetric spacetime, which has directly practical physics picture.
The system we study here requires that the string is very thin comparing with the black hole and the whole progress is free from any external force interference.
When studying the instability of the black string in the late time evolution, a similar construction has been found \cite{8}.
The microscopic construction of the place where the black string and black hole intersect is very interesting.
Another interesting field is the similarity between the black funnels and the black string flow \cite{1,2,7,9,10,21,20,13,15,16}.
Furthermore, there should be some relations between the rotating black string flow and some kinds of black funnels.
Using this method we can extend to other situations, for example, a free falling string that is not rotated falls into a rotating black hole.
Under this construction we may expect the rotation velocity of the black hole decreases to zero after the string pour enough mass to this system. Therefore, a lot of relevant works can be done.
\section*{Acknowledgments}
We gratefully acknowledge useful discussions with Dr. Ding-Fang Zeng, Dr. Jian-Feng Wu and all the others in my institute. The work is supported by National Natural Science Foundation of China (No. 11275017 and No. 11173028).
|
\section{Introduction}
We are in the era of precision cosmology. By observing the cosmic microwave background (CMB), the Planck satellite provided data that allows us to determine base quantities in the standard model of cosmology ($\Lambda$CDM) like the energy density of cold dark matter (CDM) on the percent level~\cite{Ade:2013zuv}.
Especially, if combined with further cosmological observations, most violations of assumptions and deviations from predictions are constrained severely.
We focus on recent hints for a hot dark matter (HDM) \textit{admixture}~\cite{Wyman:2013lza,Hamann:2013iba,Battye:2013xqa}
parametrised by an effective number of additional neutrino species $\DN_\text{eff}$ and effective HDM mass of~\cite{Hamann:2013iba}
\begin{equation}
\label{hdmsignal}
\DN_\text{eff} = 0.61\pm0.30 \text{,} \quad m_\text{hdm}^\text{eff} = (0.41 \pm 0.13) \text{ eV} \, .
\end{equation}
Only recently, precisions became high enough to possibly find evidence for such a \textit{sub-eV, not fully-thermalised} species. Consequently, it is of utmost importance to scrutinise and improve the accuracy of the considered observations, see for example~\cite{Costanzi:2013bha,Efstathiou:2013via,Paranjape:2014lga,Leistedt:2014sia}. For the same reason, the time to consider such HDM admixtures independent of the exact confidence intervals is now.
The common particle physics interpretation is an additional, uncharged (= sterile) neutrino $\ensuremath{\nu_s}$ species that mixes with the active neutrinos and, consequently, is produced when these scatter in the early universe.
The beauty of this interpretation is its parsimony: Firstly, there are hints from neutrino oscillation experiments for sterile neutrinos with \order{\text{eV}} masses and mixings that thermalise them. No further physics needs to be introduced, actually, from symmetry arguments one might expect even three such particles. Secondly, the cosmological model is amended by only one free parameter per particle, the sterile neutrino's mass, since full-thermalisation fixes its temperature to the neutrino temperature and thus implies $\DN_\text{eff}=1$.
However, the cosmological signal just does not fit that interpretation as illustrated in Fig.~\ref{fig:posterior}.
\begin{figure}%
\centering
\includegraphics[width=0.8\columnwidth]{dof.pdf}
\caption{Our nomenclature for a cosmological particle decay and the corresponding degrees of freedom.
The, in principle, arbitrarily large number is argued to reduce to three, which is further reduced to only two actually independent degrees of freedom by a physical parameter degeneracy.}
\label{fig:dof}
\end{figure}
Accepting the possibility of $\DN_\text{eff} < 1$, the cosmological model is amended by a second free parameter.
Attempts to reconcile sterile neutrinos with cosmology include the addition of light, interacting particles~\cite{Steigman:2013yua} and new neutrino interactions~\cite{Hannestad:2013ana,Dasgupta:2013zpn} that, indeed, lead to HDM improving the cosmological fit if they are gauged and shared by CDM~\cite{Bringmann:2013vra}. Along the way, these ideas also give up the first parsimony argument.
At that state we find it to be indicated to reconsider the interpretation of~\eqref{hdmsignal} being due to a thermally produced sterile neutrino (\ensuremath{\nu_s} HDM), even if for the sake of identifying the limits of our understanding.
In this work, we consider the possibility that hot dark matter is formed by the decay products of an out-of-equilibrium particle decay (dpHDM). So it originates from particle decay instead of scattering.
Cosmological particle decays with in part dramatic consequences are naturally expected in many theories beyond the Standard Model of particle physics (SM). They do not only occur naturally, if gravity with the prominent gravitino and Polonyi problems is considered, see~\cite{Hasenkamp:2011xh,Hasenkamp:2011em,Hasenkamp:2012ii,Bae:2013qr,Graf:2013xpe,Higaki:2013vuv,DiBari:2013dna,Conlon:2013isa,Park:2013bza,Hooper:2013nia,Kelso:2013nwa,Jeong:2013oza} and references therein.
It is remarkable that dpHDM does not require the introduction of any small mass scales as \ensuremath{\nu_s} HDM.
From a particle physical point of view a decay, as depicted schematically in Fig.~\ref{fig:dof}, might be described by an arbitrarily large number of degrees of freedom just like a ``dark sector'' with new interactions. As in that case the HDM component is, nevertheless, described by only two actually independent degrees of freedom. So from an observational point of view the complexity of the cosmological models is the same.
We will explore, which scenarios of dpHDM have for what reason the potential to be distinguished from \ensuremath{\nu_s} HDM and in which case dpHDM
is indistinguishable from \ensuremath{\nu_s} HDM in analyses like~\cite{Hamann:2013iba} utilising CMB and large-scale structure observations.
In any case, the cosmological impact of dpHDM cannot be identical, because dpHDM possesses a different momentum distribution function than \ensuremath{\nu_s} HDM. We will show at which point and how \textit{any mimicry must break down}.
This work is organised as follows:
In the next section we list main physical effects, define parameters and remind the case of \ensuremath{\nu_s} HDM, where we also assign briefly the current observational evidence.
In Sec.~\ref{sec:mimicry} we show how decay products mimic sterile neutrinos, translate observations, point out in which cases there are testable differences in the main physical effects, and how any mimicry must break down in the third cosmological parameter.
In Sec.~\ref{sec:conclusions} we summarise and conclude.
\section{Preliminaries}
\label{sec:preliminaries}
{\it \bf Physical effects (neutrino case) --}
A population of free-streaming particles, which becomes non-relativistic
after photon decoupling, affects the cosmological background and the evolution of perturbations~\cite{Lesgourgues:2006nd}.
Its main physical effects are related to:
1) its contribution to the radiation energy density $\rho_\text{rad}$ of the Universe before photon decoupling.
Given
in terms of an effective number of neutrino species $N_\text{eff}$ the radiation energy density reads
\begin{equation}
\label{rhorad}
\rho_\text{rad} = \left( 1 + N_\text{eff} \frac{7}{8} \left(\frac{T_\nu}{T_\gamma}\right)^4 \right) \rho_\gamma \, ,
\end{equation}
such that it is split into a sum of the
energy density in photons $\rho_\gamma=(\pi^2/15) T^4$ and the relativistic energy density in anything else.
The Standard Model of particle physics (SM) contains three active neutrinos with $N_\text{eff}^\text{sm}=3.046$~\cite{Mangano:2005cc}
and temperature ratio $T_\nu/T_\gamma = (4/11)^{1/3}$.
The small correction in $N_\text{eff}^\text{sm}$ is due to incomplete neutrino decoupling at $e^+e^-$-annihilation.
Any departure from the standard scenario, which increases the expansion rate of the Universe and could shift the time of matter-radiation equality, is then parametrised as a summand in $N_\text{eff} = N_\text{eff}^\text{sm} + \DN_\text{eff}$,
such that the active neutrinos correspond to $\DN_\text{eff}=0$ by construction.
2) its non-relativistic energy density today $\rho_\text{hdm}^0=n_\text{hdm}^0 m_\text{hdm}$,
which is given by the mass $m_\text{hdm}$ and today's number density $n_\text{hdm}^0$ of the population.
Free-streaming particles do not cluster on scales below the free-streaming scale and thus damp fluctuations.
The extent of the arising amplitude reduction in the matter power spectrum due to the free-streaming population is $\simeq 8 \,\Omega_\text{hdm}/\Omega_\text{m}$~\cite{Lesgourgues:2006nd}, so that the HDM fraction $f=\Omega_\text{hdm}/\Omega_\text{m}$ provides a useful parametrisation. Note that such a small HDM \textit{admixture} leads to a step-like feature and not, for example, to a cut-off.
Assuming the active neutrinos are degenerate in mass, their today's energy density normalised by the critical energy density $\rho_{\text{c}}$ reads
\begin{equation}
\label{rhonu}
\Omega_\nu^0 =
\frac{N_\text{eff}^\text{sm}}{11} \frac{n_\gamma^0}{\rho_{\text{c}}} \sum m_\nu \Leftrightarrow \Omega_\nu^0 h^2 \simeq 0.0108 \frac{\sum m_\nu}{\text{eV}}
\end{equation}
with today's number density of CMB photons given by $n_\gamma^0 = (2 \zeta(3)/\pi^2) T_0^3$,
if $T_0=2.7255$ K denotes the CMB temperature today and $\zeta$ the zeta function.
By construction~\eqref{rhonu} corresponds to the $\DN_\text{eff} = 0$ case.
In a $\Lambda$CDM model amended to include massive (degenerate) active neutrinos this sum of neutrino masses $\sum m_\nu$ is usually ``observed,''
when cosmological parameters are determined.
If $\sum m_\nu$ is not freed in an analysis (``pure'' $\Lambda$CDM), a minimal-mass normal hierarchy can be assumed, which is accurately approximated for current cosmological data as a single massive eigenstate with $m_\nu = 0.06 \text{ eV}$~\cite{Ade:2013zuv}.
3) the minimum of the comoving free-streaming wavenumber $k^\text{nr}$.
This is the scale at which the suppression of fluctuations in the matter power spectrum sets in. It is given by the time of the transition when the population becomes non-relativistic~\cite{Lesgourgues:2006nd}.
In Appendix~\ref{appendix:A} we make this point explicit arriving at
\begin{equation}
\label{knr}
k^\text{nr} \simeq 4.08 \times 10^{-4} \, \Omega_\text{m}^\frac{1}{2} \fb{T^\text{nr}}{T_0}{\frac{1}{2}} h \text{ Mpc}^{-1} ,
\end{equation}
which depends on the properties of the free-streaming population via the temperature of the Universe at transition $T^\text{nr}$ only.
It is defined by
\begin{equation}
\label{defTnr}
\langle p \rangle(T^\text{nr}) = m \, ,
\end{equation}
where the angle brackets indicate the average of the particles absolute momenta, $p=|\vec p|$.
If we kept the directional information, the population is on average at rest due to the isotropy of the Universe, $\langle \vec p \rangle =0$.
The root-mean-square momentum of the population equals the mean of its distribution of absolute momenta, $\langle \vec p \rangle_\text{rms} = \langle p \rangle$, which is sometimes just called ''mean of the distribution``.
For degenerate neutrinos~\eqref{defTnr} reads
\begin{equation}
\label{Tnrnu1}
\langle p_\nu \rangle (T^\text{nr}_\nu)
= \frac{7 \pi^4}{180 \zeta(3)} \left(\frac{4}{11}\right)^\frac{1}{3} T^\text{nr}_\nu
= \frac{1}{3} \sum m_\nu
\end{equation}
giving
\begin{equation}
\label{Tnrnu2}
T^\text{nr}_\nu = \frac{1}{3} \fb{11}{4}{\frac{1}{3}} \frac{180 \zeta(3)}{7 \pi^4} \sum m_\nu \simeq 0.148 \sum m_\nu \, .
\end{equation}
We will use $T^\text{nr}$ later on when comparing different origins of HDM.
Insertion in~\eqref{knr} yields the expected result
\begin{equation}
k^\text{nr}_\nu \simeq 0.0103 \, \Omega_\text{m}^\frac{1}{2} \fb{\sum m_\nu}{\text{eV}}{\frac{1}{2}} h \text{ Mpc}^{-1} \, .
\end{equation}
The corresponding free-streaming scale evaluated today reads
\begin{equation}
\lambda^\text{fs}_\nu \simeq 24 \, \frac{\text{eV}}{\sum m_\nu} h^{-1} \text{ Mpc} .
\end{equation}
Altogether, we see that $\sum m_\nu$ sets both $\Omega_\nu^0$ (thus $f_\nu$) and $T^\text{nr}_\nu$ (thus $k^\text{nr}_\nu$). At the same time,
the neutrino temperature is fixed by SM weak interactions.
{\bf Sterile neutrino (and current evidence) --}
\begin{figure}
\includegraphics[width=\columnwidth]{Tnr.pdf}
\caption{Joint 68\%- and 95\%-credible contours of the marginalised posterior found in~\cite{Hamann:2013iba} for $\DN_\text{eff}$ and $m_\text{hdm}^\text{eff}$ of a thermal sterile neutrino model with their 1-d means marked by a dot. Also marked is a best-fit value and the corresponding 3-$\sigma$ range from oscillation anomalies~\cite{Giunti:2013aea} in a model with one sterile neutrino. The 1+3+1 mass scheme with two sterile neutrinos, which actually might be preferred depending on the datasets considered, implies a large minimal mass like $3.2\text{ eV}$~\cite{Kopp:2013vaa}.
}
\label{fig:posterior}
\end{figure}
As prime example for hot thermal relics we consider a light sterile neutrino $\nu_s$ that is
thermalised but with a different temperature than the active neutrinos (\ensuremath{\nu_s} HDM).\footnote{
In other scenarios sterile neutrinos might be produced non-thermally. As we are going to compare the decay case with thermal production, we do not include such possibilities and consider the sterile neutrino to possess a Fermi-Dirac distribution, see also Sec.~\ref{sec:breakdown}.
}
In this case the parameters under consideration are given by:
1) If $T_{\nu_s}$ denotes the temperature of the sterile neutrino population,
\begin{equation}
\label{DNeffnus}
\DN_\text{eff} = \left(\frac{T_{\nu_s}}{T_\nu}\right)^4 \, .
\end{equation}
Inspecting~\eqref{rhorad} we see that~\eqref{DNeffnus} holds by construction for a fermion. Thus a fully-thermalised neutrino species with the energy density $\rho_{1\nu}=(7/8) (4/11)^{4/3} \rho_\gamma$ corresponds to $\DN_\text{eff}=1$.
Other particle natures can be considered by appropriate
factors in~\eqref{DNeffnus}, for example, $4/7$ for a Nambu-Goldstone boson
or $8/7$ for a massless $U(1)$ gauge boson.
If the sterile neutrino shared the bath temperature once, but decoupled earlier than the active neutrinos,
there is a one-to-one correspondence between its decoupling temperature $T_{\nu_s}^\text{dpl}$ and its temperature
at later times. It is colder as it missed heating compared to the bath, when
particles annihilated away. So $\DN_\text{eff} = (g_\ast(T_{\nu}^\text{dpl})/g_\ast(T_{\nu_s}^\text{dpl}))^{4/3}$,
where $g_\ast$ is the effective number of relativistic degrees of freedom and the decoupling temperature depends on the sterile neutrino's couplings to the bath.
The evidence in~\cite{Hamann:2013iba} for $\DN_\text{eff} > 0$ is mainly driven by local measurements of $H_0$ and seem to be supported by lensing observations~\cite{Hamann:2013iba,Battye:2013xqa}.
Independent evidence arises from the large tensor-to-scalar ratio reported by the BICEP collaboration~\cite{Ade:2014xna}.
In the simplest model of inflation a large ratio implies a large scalar spectral index that increases the power, in particular, at higher multipoles.
Whereas an increased $N_\text{eff}$ suppresses power at higher multipoles~\cite{Hou:2012xq}, which is due to increased Silk damping caused by the increased expansion rate~\cite{Hou:2011ec}.
2) The sterile neutrino's non-relativistic energy density today,
\begin{equation}
\label{Onus}
\Omega_{\nu_s}^0 = \frac{3}{11} \frac{n_\gamma^0}{\rho_{\text{c}}} \fb{T_{\nu_s}}{T_\nu}{3} \frac{m_{\nu_s}}{\text{eV}}
\Leftrightarrow \Omega_{\nu_s}^0 h^2 \simeq 0.0106 \frac{m_\text{hdm}^\text{eff}}{\text{eV}} \, ,
\end{equation}
can be understood by comparison with~\eqref{rhonu}.
The case of interest is one sterile neutrino, which is much heavier than the active neutrinos. Thus
the sum of neutrino masses is to be replaced by the sterile neutrino mass $m_{\nu_s}$.
The active neutrinos are then taken into account separately as a single massive eigenstate with $m_\nu = 0.06 \text{ eV}$.
Number densities decrease as $\propto T^3$, which explains the temperature ratio to the third power.
We just defined the effective hot dark matter mass
\begin{equation}
m_\text{hdm}^\text{eff} = m_{\nu_s} (T_{\nu_s}/T_\nu)^{1/3} .
\end{equation}
Other particle natures can be considered by appropriate
factors in~\eqref{Onus}, for example, $2/3$ for a Nambu-Goldstone boson
or $4/3$ for a massive $U(1)$ gauge boson.
Finally, if a hot relic decouples earlier than the active neutrinos, it does
not receive any corrections from $e^+e^-$ annihilation. Since this is the case of interest, the pre-factor is smaller by $3/N_\text{eff}^\text{sm}$.
There is no factor of $1/3$, because the sterile neutrino possesses three times the mass of one
degenerate neutrino in~\eqref{rhonu}.
The current evidence for $m_\text{hdm}^\text{eff} > 0$ is mainly driven by galaxy cluster data~\cite{Wyman:2013lza,Hamann:2013iba} and supported by galaxy~\cite{Hamann:2013iba,Battye:2013xqa} as well as CMB lensing data~\cite{Battye:2013xqa}. Even before the Planck cluster count became available, cluster data has pushed evidence for $m_\text{hdm}^\text{eff} > 0$~\cite{Hou:2012xq,Burenin:2013wg}.
Corresponding HDM fractions $f$ are large enough to suppress the amplitude of matter fluctuations measured in $\sigma_8$ (= root-mean-square fluctuation in total matter in $8 \, h^{-1} \text{ Mpc}$ spheres at $z=0$, computed in linear theory) down to values inferred from these ``local'' observations.
At the same time they appear small enough not to spoil the CMB fit, see also the next paragraph.
3) The average momentum of a sterile neutrino population is lowered (or increased) compared to
the active neutrinos by their temperature ratio $T_{\nu_s}/T_\nu$.
Comparing with~\eqref{Tnrnu1} and~\eqref{Tnrnu2}, we find
\begin{equation}
\label{Tnrnus}
T^\text{nr}_{\nu_s} = \frac{180 \zeta(3)}{7 \pi^4} \fb{11}{4}{\frac{1}{3}} \frac{T_{\nu}}{T_{\nu_s}} m_{\nu_s}
\simeq 0.445 \, \frac{m_\text{hdm}^\text{eff}}{\DN_\text{eff}}
\end{equation}
Now, there is a factor of three for
the reason given two paragraphs above.
For completeness, we insert into~\eqref{knr} obtaining
\begin{equation}
k^\text{nr}_{\ensuremath{\nu_s}} \simeq 0.0178 \, \Omega_\text{m}^\frac{1}{2} \fb{m_\text{hdm}^\text{eff}}{\text{eV}}{\frac{1}{2}} \DN_\text{eff}^{-\frac{1}{2}} h \text{ Mpc}^{-1}
\end{equation}
and
\begin{equation}
\lambda^\text{fs}_{\ensuremath{\nu_s}} \simeq 8.1 \, \Omega_\text{m}^\frac{1}{2} \DN_\text{eff} \frac{\text{eV}}{m_\text{hdm}^\text{eff}} h^{-1} \text{ Mpc} .
\end{equation}
In Fig.~\ref{fig:posterior} we re-plotted the countours found in~\cite{Hamann:2013iba}.
They lie roughly around the line $T^\text{nr}_{\ensuremath{\nu_s}} = T_\gamma^\text{dpl}$ with an upper bound on $\DN_\text{eff}$.
If a population becomes non-relativistic before photon decoupling, $T^\text{nr} > T_\gamma^\text{dpl}$,
it affects the CMB directly, see e.g.~\cite{Dodelson:1995es}.
The non-observation of such an impact constrains the size of $m_\text{hdm}^\text{eff}$, in particular, for large $\DN_\text{eff}$. For smaller $\DN_\text{eff}$ the impact on the CMB decreases accordingly which allows for somewhat larger $T^\text{nr}_{\ensuremath{\nu_s}}$. However, the observations appear to be matched for the obtained best-fits.
Even though $T^\text{nr}_\nu$ and $T^\text{nr}_{\ensuremath{\nu_s}}$ differ, there is neither preference in the data for $\sum m_\nu >0$ with an additional source for ``dark radiation'' implying $\DN_\text{eff}>0$ nor for \ensuremath{\nu_s} HDM.
As only difference to HDM, dark radiation is still relativistic today.
For such cosmologies parameters are obtained from our work in the limit $m_\text{hdm}\text{, }m_1\text{, }m_2 \rightarrow 0$.
\section{Perfect(?) Mimicry}
\label{sec:mimicry}
We consider the two-body decay of a non-relativistic particle (matter or dust in the cosmological sense).
We distinguish between ``early'' cosmological decays occurring before the onset of the BBN era, so at a time $\tau<0.05 \text{ s}$, and ``late'' decays occurring during or after the BBN era with $\tau>0.05 \text{ s}$.
A decay during the CMB era deserves a dedicated study investigating
its impact on the CMB power spectrum. Therefore, we restrict ourselves to decay
times $\tau \lesssim 5.2 \times 10^{10} \text{ s}$. This is before the first
observable modes of the CMB enter the horizon~\cite{Fischler:2010xz}.
In principle, a decay as drawn schematically in Fig.~\ref{fig:dof} might have to be described
by an arbitrarily large number of parameters: mass $m$ and yield $Y$ of the decaying particle (mother) and the masses
of the decay products (daughters) $m_{1}^i$, $m_{2}^i$ in each existing decay mode $i$ that sum up to the
total width $\Gamma= \sum_i \Gamma_\text{vis}^i + \sum_i \Gamma_\text{dark}^i= \tau^{-1}$, where $\Gamma_\text{vis}^i$ and $\Gamma_\text{dark}^i$ denote partial decay widths into particles with and without electromagnetic interactions, respectively.
Thus one might naively expect that there is an arbitrarily large number of possibilities for
the daughters to show up in the outlined physical effects.
In the following, we argue that, as far as the cosmological observables are concerned,
the decay can be described by a decisively smaller number of parameters
{\bf Early decay --}
The case of interest is the decay of a mother that dominates the energy density of the Universe at her decay (the opposite is covered in the next subsection) and produces HDM in its decay that does not thermalise with the SM bath.
The majority of the energy stored in the mother must be transferred to visible particles that unavoidably thermalise. The information how the mother decayed into visible particles --contained in all $\Gamma_\text{vis}^i$-- is irrelevant. Instead, the following cosmology will depend on branching ratios $B_i$ into hot dark matter $B_\text{hdm}$ and SM particles $B_\text{vis}$ only as they fix how the mother's energy is distributed. We just argued $B_\text{vis}\simeq 1 \gg B_\text{hdm}$. If one allows for (additional) decay modes to produce the observed CDM, the corresponding branching ratio were restricted to be much smaller than the one into HDM, because the CDM number density needs to be much smaller. So $B_\text{vis} = 1- B_\text{hdm}$ is either exact, if there is no additional dark decay mode, or holds to a sufficient approximation.
Concerning the kinematics, in the case of mass-degenerate daughters, $m_1 =m_2$, we define $x_2 \equiv m_2/m = m_1/m$ as measure of the mass hierarchy between mother and daughters.
It might well be that the well-motivated
case of daughters with similar masses, $m_1 \lesssim m_2$, is indistinguishable in cosmological observations from the case of mass-degenerate daughters. This is a subtle case that we leave for future work.
If the daughters possess a large mass hierarchy, $m_1 \ll m_2$,
the interesting case would be that there is a heavier daughter forming (observable)
HDM and an effectively massless one not doing so and thus forming dark radiation.
In that case the number of daughters forming HDM $g_\text{hdm}=1$, in contrast to the mass-degenerate case with $g_\text{hdm}=2$.
We define $\delta \equiv ( m - m_2)/m_2$ as measure of the mass hierarchy between mother
and heavier daughter.
Altogether, we can describe the kinematics approximately using only one hierarchy measure, $x_2$ or $\delta$, depending on the case.
The mother's energy density together with her lifetime determine the temperature of the thermal bath after her decay $T_\text{rh}$ as\footnote{
We define this temperature as the temperature of the standard thermal bath at the time of decay $T(\tau)|_\text{rad-dom}$ calculated in radiation domination. This temperature is known as ``reheating'' temperature, therefore, the subscript. However, the Universe is not re-heated as at the end of inflation. It just cools more slowly during the decay period. From this point of view, the notation used in the next subsection, $T_\text{d}$, might be seen as appropriate for this case, too. Anyway, to prevent confusion we distinguish the two cases explicitly also in the notation.
}
\begin{equation}
\label{rhomotherinearlydec}
\rho|_\text{dec}= m n|_\text{dec}= \mu^{-1} \frac{\pi^2}{30} g_\ast^\text{rh} B_\text{vis}^{-1} T_\text{rh}^4 \, ,
\end{equation}
where the correction factor $\mu = \mu_\text{dom}\simeq 0.877$ considers the exponential decay law in the expanding Universe. We determined it by solving corresponding Boltzmann equations numerically in the limit of strong dominance, $\rho \gg \rho_\text{rad}$, cf.~\cite{Scherrer:1987rr}.
Superscripts at $g_\ast$ --or, if the entropy density $s$ is considered, $g_{\ast s}$-- always indicated the temperature at which the functions are evaluated.
From an, in principle, arbitrarily large number of degrees of freedom we are down to four: $T_\text{rh}$, $B_\text{vis}$ (or $B_\text{hdm}$), $x_2$ or $\delta$ and potentially $m$.
For an early decay the cosmological parameters under consideration are given by:
1) The relativistic HDM energy density $\rho_\text{hdm}|_\text{rel} = n_\text{hdm} \langle p \rangle \simeq n_\text{hdm} (\mu/2) m$, where it has been exploited that the daughters must be much lighter than the mother due to structure formation constraints~\cite{Hasenkamp:2012ii}. With $n_\text{hdm} = g_\text{hdm} n$ it can be written as $\rho_\text{hdm}|_\text{rel} = \mu_\text{dom} \rho|_\text{dec} B_\text{hdm} (g_\text{hdm}/2) (T/T_\text{rh})^4 (g_{\ast s}/g_{\ast s}^\text{rh})^{4/3}$.
Inserting~\eqref{rhomotherinearlydec} we see that
\begin{equation}
\label{DNeffearlydec}
\DN_\text{eff} = \frac{\rho_\text{hdm}|_\text{rel}}{\rho_{1\nu}} \simeq 8.67 \, \frac{B_\text{hdm}}{B_\text{vis}} \fb{g_{\ast s}^0}{g_\ast^\text{rh}}{\frac{1}{3}} \, .
\end{equation}
2) Its today's (non-relativistic) energy density $\rho_\text{hdm}^0 = n^0_\text{hdm} m_\text{hdm}$ can be written as $\rho^0_\text{hdm}= \rho|_\text{dec} x_2 g_\text{hdm} B_\text{hdm} (T_0/T_\text{rh})^3 (g_{\ast s}^0/g_{\ast s}^\text{rh})$. Inserting~\eqref{rhomotherinearlydec} we see that it reads
\begin{equation}
\label{Ohdmearlydec}
\Omega^0_\text{ed} h^2 = \frac{\rho_\text{hdm}^0 h^2}{\rho_{\text{c}}} \simeq 4.15 \, \mu_\text{dom}^{-1} \frac{x_2}{10^{-8}} \frac{B_\text{hdm}}{B_\text{vis}} \frac{T_\text{rh}}{\text{GeV}}\, .
\end{equation}
3) The temperature of the Universe when the population of free-streaming daughters becomes non-relativistic is found as
\begin{eqnarray}
T^\text{nr}_\text{ed} &=& T_\text{rh} \frac{2}{\mu} \frac{\delta+1}{(\delta+1)^2 -1} \fb{g_{\ast s}^\text{rh}}{g_{\ast s}^0}{\frac{1}{3}} \nonumber \\
\text{ or } \, &=& T_\text{rh} \frac{2}{\mu} (x_2^{-2} -4)^{-\frac{1}{2}} \fb{g_{\ast s}^\text{rh}}{g_{\ast s}^0}{\frac{1}{3}} ,
\label{Tnrdecay}
\end{eqnarray}
see (8) and (50) of~\cite{Hasenkamp:2012ii}, while $\mu = \mu_\text{dom}$, here. %
In all cases $T^\text{nr}_\text{dec}< T_\text{eq}$ and the minimal comoving free-streaming wavenumber of the daughters is set by~\eqref{knr} with
$T^\text{nr}$ given by~\eqref{Tnrdecay}.
By inspection of~\eqref{DNeffearlydec} and~\eqref{Ohdmearlydec} we see that the mother's mass does not enter independently. Furthermore, the observables depend on the ratio of the branching ratios $B_\text{hdm}/B_\text{vis}$ only. In other words, they depend on the amount of HDM relative to the amount of visible matter at the decay. This reduces the number of degrees of freedom in the description of the early decay to three: $T_\text{rh}$, $B_\text{hdm}/B_\text{vis}$ and a hierarchy measure, either $\delta$ or $x$, depending on the case under investigation.
A dominating mother produces significant entropy that dilutes relic densities during her decay. The corresponding dilution factor $\Delta$ can be given as~\cite{Hasenkamp:2010if}
\begin{equation}
\Delta = \frac{\langle g_\ast^\frac{1}{3} \rangle^\frac{3}{4}}{(g_\ast^\text{rh})^\frac{1}{4}} \frac{mY}{T_\text{rh}} \, ,
\end{equation}
where the angle brackets indicate the appropriately-averaged value of $g_\ast$ over the decay interval. We see that $\Delta$ is independent of the HDM observables. The dilution is an \textit{additional, independent} physical effect. After discovering the primordial gravitational wave background in the CMB, we \textit{know} that this impact of the decay will be ultimately tested by a local detection of this background radiation~\cite{Durrer:2011bi} that would rule out any such early period of matter domination .
{\it \bf Late decay --}
For $\tau > 0.05 \text{ s}$ current constraints on $\DN_\text{eff}$ forbid the particle to dominate the energy
density of the Universe prior its decay, see Sec.~2.1 of~\cite{Hasenkamp:2012ii}.\footnote{
Of course, the following considerations hold for any earlier decay of a non-relativistic particle that does not dominate at its decay. There is just no observational reason that would forbid its domination.
}
For such late decays, if the mother is sufficiently heavier
than the proton, the branching ratio into any visible particles is constrained
from BBN and CMB observations to be much smaller
than one, $\sum_i \Gamma_\text{vis}^i \ll \sum_i \Gamma_\text{dark}^i$, see Sec.~3 of~\cite{Hasenkamp:2012ii}.
This holds even if photons and electrons are emitted at the end of a decay chain only, cp.~\cite{Cline:2013fm}.
If one allows for (additional) decay modes to produce the observed dark matter,
the corresponding branching ratio were restricted to be much smaller than the one into
hot dark matter, $\sum_i \Gamma_\text{dark}^i \simeq \Gamma_\text{hdm}$, see Sec.~4.2 of~\cite{Hasenkamp:2012ii}.
In other words, the HDM branching ratio is to a very good approximation one
and the lifetime of the mother is given by her HDM decay width.
This implies a decisive reduction of parameters. From an, in principle, arbitrarily large number
we are down to four: mass $m$, yield $Y$ and lifetime $\tau$ of the mother and one of the two hierarchy measures, $\delta$ or $x_2$.
For the late decay following~\cite{Hasenkamp:2012ii}
the cosmological parameters under consideration are given by:
1) If $T_\text{d}= T(\tau)$ denotes the
temperature of the Universe at decay\footnote{
The Universe is radiation dominated in the time window under consideration.
},
\begin{equation}
\label{DNeffdecay}
\DN_\text{eff} = 10.25 \, \frac{m Y}{T_\text{d}} \fb{g_{\ast s}^0}{g_{\ast s}^\text{d}}{\frac{1}{3}} \frac{(\delta+1)^2-1}{(\delta+1)^2} \, .
\end{equation}
Again taking into account the huge mass hierarchy, $\delta\gg 1$, required from structure formation constraints,
the last fraction is also in this case to a good approximation one. This holds analogously, if the daughters are mass degenerate, see~(20) of~\cite{Hasenkamp:2012ii}.
2) The non-relativistic energy density of the daughters today is given by, cf.~(45) and~(55) of~\cite{Hasenkamp:2012ii},
\begin{align}
\label{Odecay}
\Omega_\text{ld}^0 &= g_\text{hdm} \frac{mY}{\delta +1} \frac{g_{\ast s}^\text{d}}{g_{\ast s}^0} \frac{s_0}{\rho_{\text{c}}} \nonumber \\
& \Leftrightarrow \Omega_\text{ld}^0 h^2 = \frac{2.76 \times 10^8}{\delta+1} \frac{mY}{\text{GeV}} g_\text{hdm} \frac{g_{\ast s}^\text{d}}{g_{\ast s}^0}\, .
\end{align}
For a decay into mass-degenerate daughters we should replace $1/(\delta +1) \rightarrow x_2 $.
Note that the yield $Y$ is the one in~\eqref{DNeffdecay}. So it is to be evaluated at the mother's decay.
nto two identical particles we should replace $1/(\delta +1) \rightarrow 2 x_2 = 2 m_2 /m$.
Note that by construction the same decay
leads to both, $\DN_\text{eff}>0$ and $m_\text{hdm}^\text{eff}>0$. We discuss the implications of a second
decay mode into dark radiation or HDM below.
3) The temperature of the Universe when the population of free-streaming daughters becomes non-relativistic is given by~\eqref{Tnrdecay} with $\mu = \sqrt{\pi}/2$.
We inspect~\eqref{DNeffdecay} and~\eqref{Odecay} to complete our counting of degrees of freedom in the description of a late cosmological particle decay, for the moment. We see that both observables
depend on the product $mY$ only and not on $m$ or $Y$ independently. They depend on the energy
density at decay and not on how the energy density is built from a large mass or number density, as for an early decay.
Thus observations are not sensitive to this degeneracy in determining the energy density of the
mother.
So instead of an arbitrarily large number of degrees of freedom
we identify three: the lifetime $\tau$, the energy density of the decaying particle $\rho(=mYs)$
and a hierarchy measure, either $\delta$ or $x$, depending on the case under investigation.
Even though the descriptions for the early and late decay are qualitatively different,
we find in both cases three degrees of freedom that are possibly affecting the outlined three main physical effects of a
free-streaming population. One might compare this to \ensuremath{\nu_s} HDM with
two parameters, i.e., $T_{\nu_s}$ and $m_{\nu_s}$.
In contrast to \ensuremath{\nu_s} HDM, no (possibly unnatural) small mass scale needs to be introduced.
\subsection{Translating observations}
\begin{figure}
\includegraphics[width=\columnwidth]{DNeffdecay.pdf}
\label{fig:DNeffdecay}
\caption[{$mY(\tau)$ fixed by $T_{\nu_s}$}]{ Determination of $mY(\tau)$ depending on $T_{\nu_s}$ exploiting~\eqref{DNeffdecay}.
A higher $T_{\nu_s}$ is mimicked by a correspondingly larger energy density of the decaying particle at decay.
The solid line corresponds to $N_\text{eff} = 0.61 \pm 0.60$ with its 2-$\sigma$ range as grey band.
The dashed line marks Planck's CMB only mean $\DN_\text{eff}=0.29$~\cite{Ade:2013zuv}.
}
\end{figure}
\begin{figure}%
\centering
\includegraphics[width=\columnwidth]{Odecay.pdf}
\label{fig:Odecay}
\caption{Determination of $mY(\delta)$ depending only on $m_\text{hdm}^\text{eff}$.
A larger $m_\text{hdm}^\text{eff}$ is mimicked by a correspondingly smaller mass hierarchy.
The solid line corresponds to $m_\text{hdm}^\text{eff}/\text{eV} = 0.41 \pm 0.26$ with its 2-$\sigma$ range as grey band.
The dashed line marks Planck's CMB-only upper bound $m_\text{hdm}^\text{eff} < 0.59 \text{ eV}$ (95\%)~\cite{Ade:2013zuv}.
We have chosen $g_\text{hdm}=1$ for the figure. If the decay is into mass-degenerate daughters, the horizontal axis should carry $x_2^{-1}$ and all lines shift down by $1/2$ as $g_\text{hdm}=2$.
}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{Odecay2.pdf}
\label{fig:Odecay2}
\caption{Determination of $\delta(\tau)$ depending only on $m_\text{hdm}^\text{eff}$
exploiting~\eqref{DNeffdecay} and~\eqref{Odecay} with fixed $T_{\nu_s}$.
For a fixed $m_\text{hdm}^\text{eff}$ the later the decay the smaller the mass hierarchy might be.
The solid line corresponds to $m_\text{hdm}^\text{eff}/\text{eV} = 0.41 \pm 0.26$ with its 2-$\sigma$ range as grey band, while $\DN_\text{eff}=0.61$ is kept fix.
The jump around $t\sim 10 \text{ s}$ corresponds to the time of $e^+e^-$ annihilation.
Planck's CMB only upper bound $m_\text{hdm}^\text{eff} < 0.59 \text{ eV}$ and $\DN_\text{eff} < 0.86$ (both 95\%)~\cite{Ade:2013zuv} is barely visible right on the solid line.
We have chosen $g_\text{hdm}=1$ for the figure. If the decay is into identical particles, the vertical axis should carry $x_2^{-1}$ and all lines shift up by $1/2$ as $g_\text{hdm}=2$.
}
\end{figure}
In the following we show actually that dpHDM and \ensuremath{\nu_s} HDM are in general indistinguishable in CMB and large-scale structure observations using $N_\text{eff}$ and $m_\text{hdm}^\text{eff}$ only.
First, we provide the prescription how to obtain
(for any ``observed'' sterile neutrino parameter)
the corresponding
parameters describing the cosmological particle decay
with exactly the same signal.
In other words, parameters determined assuming a sterile neutrino cosmology
can be translated back and forth into parameters describing a cosmology with
HDM originating from a cosmological particle decay.
{\bf Early decay --}
Comparing~\eqref{DNeffnus} and~\eqref{DNeffearlydec} we see trivially that a higher sterile neutrino temperature is mimicked by a larger $B_\text{hdm}/B_\text{vis}$ (for fix $g_\ast^\text{rh}$).
Furthermore, equating~\eqref{Onus} and~\eqref{Ohdmearlydec} we find how the effective \ensuremath{\nu_s} HDM mass is mimicked,
\begin{equation}
\label{mnuseffearlydec}
m_\text{hdm}^\text{eff} = \mu_\text{dom}^{-1} \frac{11 \pi^4 g_{\ast s}^0}{90 \zeta(3)} \frac{B_\text{hdm}}{B_\text{vis}} x_2 T_\text{rh} \, .
\end{equation}
So the prescription is: i) Fix $B_\text{hdm}/B_\text{vis}$ to obtain the desired $\DN_\text{eff}$. ii) Determine $x_2 T_\text{rh}$, while $\DN_\text{eff}$ is kept fix. At this point keep in mind that the description is valid for $T_\text{rh} > \text{ MeV}$ only.
For illustration, we translate~\eqref{hdmsignal} into
\begin{equation}
\label{sigDNeffearlydec}
\frac{B_\text{hdm}}{B_\text{vis}} \left( g_\ast^\text{rh}\right)^{-\frac{1}{3}} = 0.045 \pm 0.022
\text{ } (\text{``1-}\sigma\text{''} \text{; } \tau < 0.05 \text{ s})
\end{equation}
and
\begin{align}
\label{sigmnuseffearlydec}
\frac{B_\text{hdm}}{B_\text{vis}} \frac{x_2}{10^{-9}} \frac{T_\text{rh}}{\text{GeV}} = (9.2 \pm 2.9) & \times 10^{-3} \nonumber \\
& (\text{``1-}\sigma \text{''; } \tau < 0.05 \text{ s})
\end{align}
for a scenario with mass-degenerate daughters.
If the daughters possess a large mass hierarchy, such that $g_\text{hdm}=1$,
a) nothing changes for $\DN_\text{eff}$ and
b) $x_2 \rightarrow 1/(\delta+1)$ in~\eqref{sigmnuseffearlydec} and its right-hand-side is to be divided by 2.
However, in this case there is a shift in $T^\text{nr}$ as shown below and thus a difference in the main physical effects.
We would like to point out that~\eqref{sigDNeffearlydec} and~\eqref{sigmnuseffearlydec} (later also~\eqref{sigDNeff} and~\eqref{sigmnuseff}) are by way of illustration only. We simply ``translated'' parameters, while the confidence intervals are not determined by a likelihood ratio, but by the area below the posterior function that is not invariant under non-linear parameter transformations.
In Sec.~\ref{sec:breakdown} we identify (two, independent) parameters that should be determined in a Markov Chain Monte Carlo likelihood analysis with approriate priors.
{\bf Late decay --}
How to mimic a sterile neutrino in $N_\text{eff}$ is shown in Fig.~\ref{fig:DNeffdecay}.
The horizontal axis spans the full range of considered lifetimes.
We take the temperature dependence of $g_{\ast s}$ fully into account.
Around $\tau\sim10 \text{ s}$, when $e^+e^-$ annihilate away, $g_{\ast s}$ decreases. The curves stay straight as for the same reason the temperature of the Universe increases, both effects cancelling each other up to the difference between $g_\ast^0\simeq3.384$ and $g_{\ast s}^0\simeq3.938$. It is roughly $(g_{\ast s}^0/g_\ast^0)^{1/3} \simeq 1.05$.
The time when BBN ends $t_\text{bbn}^\text{end}$ is highlighted.
Decays before this time increase the effective number of neutrino species inferred from
BBN $\DN_\text{eff}|_\text{bbn}$ partially as quantified in~\cite{Menestrina:2011mz}.
Comparing~\eqref{DNeffnus} and~\eqref{DNeffdecay} we see that a higher $T_{\nu_s}$
is mimicked by a larger energy density of the mother at decay.
This has been shown in~\cite{Hooper:2011aj} solving the corresponding Boltzmann equations numerically.
The values chosen for $\DN_\text{eff}$, which are in one-to-one correspondence with $T_{\nu_s}$ via~\eqref{DNeffnus},
are given in the figure caption.
We can see how a narrow range of $T_{\nu_s}$ maps onto a narrow range of $mY(\tau)$ forming a strip in the $mY$-$\tau$ plane.
After having fixed $mY/T_\text{d}$ we can use Fig.~\ref{fig:Odecay}, which is obtained by setting~\eqref{Onus} equal to~\eqref{Odecay}, to determine the correct $\delta(mY)$
mimicking the effective HDM mass
\begin{align}
&m_\text{hdm}^\text{eff} \simeq 26.0 \, g_\text{hdm} \frac{mY}{\delta+1} \frac{g_{\ast s}^\text{d}}{g_{\ast s}^0} \nonumber \\
&\text{ or } \simeq 52.0 \, x_2 mY \frac{g_{\ast s}^\text{d}}{g_{\ast s}^0} \, .
\end{align}
Alternatively, we can use~\eqref{DNeffdecay} to re-write~\eqref{Odecay}, which equals~\eqref{Onus}, obtaining
\begin{equation}
\label{deltaofobs}
\delta \simeq 5 \times 10^3 \, \frac{T_\text{d}}{\text{keV}} \DN_\text{eff} \frac{\text{eV}}{m_\text{hdm}^\text{eff}}
\fb{g_{\ast s}^\text{d}}{g_{\ast s}^0}{\frac{4}{3}} \frac{g_\text{hdm}}{2} -1
\end{equation}
which allows to provide Fig.~\ref{fig:Odecay2}.
For a given lifetime we can thus read off the corresponding $mY$ and $\delta$ to
mimic any ``observed'' sterile neutrino parameters.
Altogether, we found the following prescription:
i) Determine $mY(\tau)$ using~\eqref{DNeffdecay} such that the desired $\DN_\text{eff}$ is obtained.
ii) Determine $\delta(\tau)$ using~\eqref{deltaofobs} such that the desired $m_\text{hdm}^\text{eff}$ is obtained while $\DN_\text{eff}$ is kept fixed.
This makes dpHDM indistinguishable from \ensuremath{\nu_s} HDM
in $N_\text{eff}$ and $m_\text{hdm}^\text{eff}$.
For illustration, we translate~\eqref{hdmsignal} into
\begin{align}
\label{sigDNeff}
\frac{mY}{T_\text{d}}\fb{g_{\ast s}^0}{g_{\ast s}^\text{d}}{\frac{1}{3}} &= 0.060 \pm 0.029 \nonumber \\
& (\text{``1-}\sigma \text{''; } 0.05 \text{ s} <\tau <5.2 \times 10^{10} \text{ s})
\end{align}
and
\begin{align}
\label{sigmnuseff}
x_2 \, mY \frac{g_{\ast s}^\text{d}}{g_{\ast s}^0} &= (7.9 \pm 2.5)\times 10^{-3} \text{ eV} \nonumber \\
&(\text{``1-}\sigma \text{``; } 0.05 \text{ s} <\tau <5.2 \times 10^{10} \text{ s})
\end{align}
for a scenario with mass-degenerate daughters.
If the daughters possess a large mass hierarchy, such that $g_\text{hdm}=1$, proceed as discussed below~\eqref{sigmnuseffearlydec}.
It is important that --in contrast to the early decay case-- the assumption of BBN consistency is wrong, because for a late cosmological decay $N_\text{eff}|_\text{bbn} < N_\text{eff}|_\text{cmb}$. The possibility of late cosmological decays motivates determinations of $N_\text{eff}|_\text{cmb}$, where the primordial abundance of light elements is constrained from direct observations only, so \textit{really} independent of cosmology at earlier times, and fixed $N_\text{eff}|_\text{bbn}=N_\text{eff}^\text{sm}$. For earlier related work with a different approach in a different scenario see~\cite{GonzalezGarcia:2012yq}.
{\bf A degeneracy --}
For the early decay we can exploit~\eqref{DNeffearlydec},~\eqref{Tnrdecay} and~\eqref{mnuseffearlydec} to obtain
the following \textit{parameter degeneracy} relation
\begin{equation}
\label{Tnrobs2}
T^\text{nr}_\text{ed} = \frac{180 \zeta(3)}{7 \pi^4} \fb{11}{4}{\frac{1}{3}} \frac{m_\text{hdm}^\text{eff}}{\DN_\text{eff}} \frac{2}{g_\text{hdm}} \, .
\end{equation}
Analogously, if we use~\eqref{DNeffdecay} and~\eqref{Odecay} eliminating $mY$, we can single out
$T_\text{d} (\delta+1)/ ((\delta+1)^2 -1)$. Inserting the resulting expression into~\eqref{Tnrdecay} we obtain
\begin{equation}
\label{Tnrobs}
T^\text{nr}_\text{ld} \simeq 0.445 \frac{m_\text{hdm}^\text{eff}}{\DN_\text{eff}} \frac{g_{\ast s}^0}{g_{\ast s}^\text{d}} \frac{2}{g_\text{hdm}} \, .
\end{equation}
We see that the main physical effects are described by only two independent parameters.
In other words, there is a degeneracy among the three physical parameters.
The same degeneracy has been identified for the thermal relic case in~\cite{Acero:2008rh}.\footnote{
It has been used to show that a Dodelson-Widrow model shares the same ''observable`` parameters
as a thermal sterile neutrino model with adjusted mass and temperature.
}
Indeed, we find exactly the same degeneracy, i.e. the same numerical value, if we
consider $g_\text{hdm}=2$ and -for the late decay- $g_{\ast s}^\text{d} = g_{\ast s}^0$, which is fulfilled after $e^+e^-$ annihilation, so in a large part of the parameter space.
Viewed as a constraint equation the degeneracy in physical parameters reduces
the number of independent degrees of freedom describing dpHDM from three to two.
Considering $N_\text{eff}$ and $m_\text{hdm}^\text{eff}$ only, which can describe the
main physical effects,
an early cosmological particle decay, as well as a decay after $e^+e^-$ annihilation, as origin of
HDM is indistinguishable from \ensuremath{\nu_s} HDM, if the daughters are mass degenerate.
The mimicry is ''perfect`` in the sense that there is no new signature
in the main physical effects as those contained in \ensuremath{\nu_s} HDM parameters ($T_{\nu_s}$ or $\DN_\text{eff}$ and $m_{\nu_s}$ or $m_\text{hdm}^\text{eff}$).
{\it \bf More complex scenarios}
If the daughters do not possess a large
mass hierarchy, $m_1\lesssim m_2$, their impact on observables is similar to the case
of two sterile neutrinos with equal temperature but different mass.
Moreover, considering two sterile neutrino populations with sufficiently different temperatures, one neutrino might not account for
$\Omega_\text{hdm}$ as being still relativistic today and thus forming dark radiation. This situation would correspond to
two dark decay modes in the cosmological particle decay, one for dark radiation
and one for HDM. All we would like to point out here is that these scenarios
are more complex also in the case of a thermal relic. Consequently, they are not discussed in this work and left for future work.
The arguably most interesting case is found if the heavier daughter forms HDM and the lighter one dark radiation. Then the temperature of the universe, when the
HDM becomes non-relativistic is a factor of two larger as $g_\text{hdm}=1$ in~\eqref{Tnrobs2} and~\eqref{Tnrobs}, respectively.
Translated to sterile neutrinos, this situation is represented by two sterile neutrino populations with the same temperature, while one is massive and the other one is massless.
Interestingly, current (non-zero) mean values for a HDM contribution correspond
to $T^\text{nr}$s around $T_\gamma^\text{dpl}$, see Fig.~\ref{fig:posterior}.
If $T^\text{nr}$ becomes significantly larger than $T_\gamma^\text{dpl}$ due this factor of two,
this effect could be used to distinguish this case, because the impact on the CMB
becomes qualitatively different for $T^\text{nr} >T_\gamma^\text{dpl}$~\cite{Dodelson:1995es}.
As mentioned already, the non-observation of such an impact constrains the size of $m_\text{hdm}^\text{eff}$.
While we do not attempt quantitative statements, here, it seems that the increase in $m_\text{hdm}^\text{eff}$ would be in stronger tension with the CMB data for $g_\text{hdm}=1$. This could disfavour this case compared to the case of mass-degenerate daughters.
On the other hand, if a late decay occurs before $e^+e^-$ annihilation, the daughters become non-relativistic at a time $g_{\ast s}^\text{d}/g_{\ast s}^0 \simeq 2.73$ later which reduces the constraining power from the direct impact on the CMB.
For lower $T^\text{nr}$ it has been argued in~\cite{Hannestad:2005bt} that an optimal LSST-type wide field survey might provide the ability to distinguish
between thermal, fermionic HDM (Majorana fermion) and thermal, bosonic HDM (scalar).
For fixed $\Omega_\text{hdm}$ the scalar becomes non-relativistic at a temperature,
which is a factor of $3/2$ larger than in the case of a fermion. This is simply
the increase in mass compensating the factor considering the scalar particle nature in~\eqref{Onus}.
We might conclude that the same observations are able
to differ between \ensuremath{\nu_s} HDM and dpHDM where the heavier daughter forms HDM and the lighter one dark radiation. However, on the observed non-linear scales N-body simulations appear to become necessary~\cite{Bird:2011rb} and HDM infall might become an observable effect with the increased precision~\cite{Hannestad:2005bt}.
The exploration of this interesting opportunity to probe the origin of HDM is beyond the scope of this work.
In future work, all discussed scenarios can be implemented in CAMB\footnote{
http://camb.info/
} or CLASS\footnote{
http://class-code.net/
} easily, because they can be represented by thermal fermion populations as just described and summarised in Tab.~\ref{tab:summary}.
\subsection{Breakdown in the Third}
\label{sec:breakdown}
\begin{figure}
\includegraphics[width=\columnwidth]{P_x__overplot.pdf}
\caption{Normalised, time-invariant probability distribution $P(x=p/p(\tau,t))$ for finding a relativistic particle from two-body decay (decaying particle at rest) within the infinitesimal momentum interval $[p,p+\delta p]$. In the early decay case the mother may dominate (black, solid) the Universe. For a late decay the Universe is radiation dominated (black, dashed). The Universe becomes matter-dominated (grey, dotted) at very late times only.
Highlighted with corresponding lines are the maxima and means.
}
\label{fig:Px_overplot}
\end{figure}
It is a simplification to reduce the physical impact of any population of massive free-streaming particles
to the effects listed in Sec.~\ref{sec:preliminaries}. The free-streaming
effect must depend on the details of the momentum distribution $f(p)$ of the population.
For illustration imagine a distribution with a sharp peak close to $p=0$. Such particles would
act as cold dark matter and thus should not be counted within the massive free-streaming component~\eqref{Onus}.
Furthermore, the average considered in~\eqref{defTnr} were between cold and hot particles, so that
the correct physical effects were not captured.
{\bf Observables and statistical moments --}
Distribution functions can be specified by their moments $Q^{(n)}$.
This has been done for the Fermi-Dirac
(active neutrino) distribution $f_\text{fd}(y)=1/(e^y +1)$ with comoving momentum $y=pa$ in~\cite{Cuoco:2005qr}.
As their result can be adapted to the sterile neutrino case straightforwardly, we
sketch decisive steps in Appendix~\ref{appendix:nudistr}.
For a thermal population of sterile neutrinos we define
\begin{equation}
\label{Qnus}
Q^{(n)}_{\ensuremath{\nu_s}} = \frac{1}{\pi^2} \fb{4}{11}{\frac{3+n}{3}} \fb{T_{\nu_s}}{T_\nu}{3+n} T^{3+n}
\int y^{2+n} f_\text{fd}(y) dy \, .
\end{equation}
The (first two) cosmological parameters can be expressed by these moments as
\begin{equation}
\label{NeffbyQnus}
N_\text{eff} = 3.046 + \frac{120}{7 \pi^2} \fb{11}{4}{\frac{4}{3}} T^{-4} Q^{(1)}_{\ensuremath{\nu_s}}
\end{equation}
and
\begin{align}
\Omega_{\ensuremath{\nu_s}}^0 &= \frac{m_{\nu_s}}{\rho_{\text{c}}} \fb{T_0}{T}{3} Q_{\ensuremath{\nu_s}}^{(0)} \Leftrightarrow \nonumber \\
\Omega_{\ensuremath{\nu_s}}^0 h^2 & \simeq 0.160 \, \frac{m_\text{hdm}^\text{eff}}{\text{eV}} \DN_\text{eff}^\frac{3}{4} T^{-3} Q_{\ensuremath{\nu_s}}^{(0)} \, .
\end{align}
We see that $N_\text{eff}$ and $\Omega_{\ensuremath{\nu_s}}^0$ probe the first two moments of the momentum distribution
function.
If the mother does not dominate the Universe at her decay,
the momentum distribution function of relativistically emitted daughters (cf.\@ appendix of~\cite{Hasenkamp:2012ii} and~\cite{Scherrer:1987rr})
as function of comoving momentum $y$ reads
\begin{equation}
\label{decdistr}
df(y)= n_\text{ini} c p_\text{ini}^{-c} y^{c-1} e^{-(y/p_\text{ini})^c} dy \, ,
\end{equation}
where $n_\text{ini}$ denotes the number density of daughters at decay and $p_\text{ini}= m (1-(\delta+1)^{-2})/2$
or $=m (1 - 4 x_2^2)^{1/2}/2$ their initial momentum. For the considered lifetime range
it is well approximated by $p_\text{ini} \simeq m/2$.
In Fig.~\ref{fig:Px_overplot} we depicted the corresponding probability distribution function.
For the late decay
we use explicitly that the Universe is radiation dominated, so that $c\equiv 3(1+\omega)/2 =2$ as the equation of state
of the Universe, $p = \omega \rho$, is given by $\omega=1/3$ in that case.
Nevertheless, our treatment could be applied to any expansion law with $c>0$.
By inspection of~\eqref{decdistr} we can see that the distribution function decays as $e^{-y}$ for large comoving momenta as the
neutrino distribution function considered in~\cite{Cuoco:2005qr}.
We define the moments as
\begin{equation}
\label{Qdec}
Q_\text{dec}^{(n)} = \frac{n_\text{ini}}{T_\text{d}^3} \fb{p_\text{ini}}{T_\text{d}}{n} \fb{g_{\ast s}^0}{g_{\ast s}^\text{d}}{\frac{n}{3}} T^{3+n} \int c y^{c+n-1} e^{-y^c} dy
\end{equation}
Then the (first two) cosmological parameters expressed by these moments read
\begin{equation}
N_\text{eff} = 3.046 + \frac{120}{7\pi^2}\fb{11}{4}{\frac{4}{3}} T^{-4} Q_\text{dec}^{(1)}
\end{equation}
and
\begin{align}
\label{OQdec}
\Omega_\text{dec}^0 &= \frac{g_\text{hdm}}{2} \frac{m_2}{\rho_{\text{c}}} \fb{T_0}{T}{3} Q_\text{dec}^{(0)} \Leftrightarrow \nonumber \\
\Omega_\text{dec}^0 h^2 &\simeq 0.160 \, \frac{g_\text{hdm}}{2} \frac{m_2}{\text{eV}} T^{-3} Q_\text{dec}^{(0)}
\end{align}
Note that $n_\text{ini}= 2 n$. We identify two dimensionless, independent parameters $n_\text{ini}/T_\text{d}^3$ and $p_\text{ini}/T_\text{d}$
in accordance with the previous discussion.
In future work, these could be determined with a Markov Chain Monte Carlo likelihood analysis.
Our discussion from~\eqref{Qnus} to~\eqref{OQdec}
is an alternative way to demonstrate why the mimicry in $N_\text{eff}$ and $m_\text{hdm}^\text{eff}$ is ''perfect``.
In the case of a dominating mother, the simple analytic form~\eqref{decdistr} is invalid. Boltzmann equations need to be solved numerically, cf.~\cite{Scherrer:1987rr}. In Fig.~\ref{fig:Px_overplot} we depicted the resulting probability distribution in the limit of strong dominance, $\rho/\rho_\text{rad} \gg 1$, at decay.
Concerning the moments~\eqref{Qdec}, the factor $g_{\ast s}^0/g_{\ast s}^\text{d}$ were absent and the integral could not be performed analytically.
{\bf Root-mean-square of absolute momenta --}
The momentum distribution function of a thermal relic differs from the momentum distribution
function of relativistic decay products, cp.~\eqref{nudistr} and~\eqref{decdistr}. The two available degrees of
freedom have been used to achieve perfect mimicry in the first two cosmological parameters, which have
been expressed by the first two moments of the distribution functions.
Therefore, the \textit{mimicry must break down} if the next (third) cosmological parameter, which is to be expressed
by the next higher moment of the distribution function, is taken into account. The number
of parameters to be mimicked then exceeds the number of degrees of freedom.
We identify the third cosmological parameter as the root-mean-square of absolute momenta $\langle p \rangle_\text{rms}$ --or, equivalently, the root-mean-square of absolute velocities today\footnote{
If clustering can be neglected, today's root-mean-square of absolute velocities $\langle v^0 \rangle_\text{rms} = \langle p^0 \rangle_\text{rms}/m$.
}--, which is given by the quadratic mean of the distribution of absolute momenta.
This quadratic mean is equal to the skewness of the momentum distribution keeping the directional information.
In general, the skewness is the next higher moment following the quadratic mean. The next higher moment to the skewness is the kurtosis.
Giving preference to dimensionless parameters we define the \textit{normalised root-mean-square of absolute momenta}
\begin{equation}
\label{defnormrmsmomentum}
\gamma^\text{rms} \equiv \frac{\langle p \rangle_\text{rms}}{\langle p_\nu \rangle_\text{rms}}
\end{equation}
where $\langle p_\nu \rangle_\text{rms}$ is the root-mean-square of the absolute momentum of the active neutrinos~\eqref{rmsnumomentum}, so that $\gamma_\nu^\text{rms} = 1$ by definition.
The root-mean-squares of absolute momenta can be expressed by moments of the corresponding momentum distribution functions.
This can be understood by reminding that the number density $n = Q^{(0)}$ and the relativistic energy density $\rho =\langle p \rangle n = Q^{(1)}$, so that the mean absolute momentum $\langle p \rangle = Q^{(1)}/Q^{(0)}$. Analogously, $\langle p \rangle_\text{rms} = (Q^{(2)}/Q^{(0)})^{1/2}$.
For a thermal sterile neutrino population we find
\begin{equation}
\label{nusprms}
\langle p_{\nu_s} \rangle_\text{rms}
= \fb{15 \zeta(5)}{\zeta(3)}{\frac{1}{2}} \fb{4}{11}{\frac{1}{3}} \frac{T_{\nu_s}}{T_\nu} T \simeq 3.6 \, T_{\nu_s} \,,
\end{equation}
which implies with~\eqref{rmsnumomentum} a normalised root-mean-square of absolute momenta
\begin{equation}
\label{grmsnus}
\gamma_{\nu_s}^\text{rms} = \frac{T_{\nu_s}}{T_\nu} = \DN_\text{eff}^{\frac{1}{4}}
\end{equation}
that can be expressed by lower momenta~\eqref{NeffbyQnus}. This is to be expected, since $\gamma_{\nu_s}^\text{rms}$ cannot carry additional information as both neutrino species possess a Fermi-Dirac distribution.
The momentum distribution function of the sterile neutrinos in the popular Dodelson-Widrow (DW) scenario~\cite{Dodelson:1993je} (a.k.a.\@ ''non-resonant production scenario``) reads
\begin{equation}
f_\text{dw}(y) = \frac{\chi}{e^y+1} = \chi f_\text{fd} \,.
\end{equation}
The sterile neutrinos share the same ''observable`` parameters ($\DN_\text{eff}$, $\Omega_\text{hdm}^0$, $T^\text{nr}$) as a thermal population with $m_{\nu_s}^\text{th} = \chi^{1/4} m_{\nu_s}^\text{dw}$ and $T_{\nu_s}^\text{th} =\chi^{1/4} T_\nu$.
For these two models it has been shown by a change of variable in the background and linear perturbation equations that the two models are strictly equivalent for cosmological observables~\cite{Colombi:1995ze}.
We confirm this result with the expansion of the distribution functions in moments. Compensating the scaling factor $\chi$, the two models share the same (Fermi-Dirac) distribution function. Higher moments can be expressed by lower ones and the distribution functions are identical. Therefore, they are indistinguishable for cosmological observations even if higher moments are taken into account. In other words, their \textit{mimicry is perfect to arbitrary order in moments}.
This is \textit{qualitatively different for the decay products} of a cosmological particle decay mimicking a thermal sterile neutrino population.
For daughters emitted in a late decay we find
\begin{equation}
\langle p_\text{dec} \rangle_\text{rms} =
\fb{Q_\text{dec}^{(2)}}{Q_\text{dec}^{(0)}}{\frac{1}{2}}
= \frac{p_\text{ini}}{T_\text{d}} T \fb{g_{\ast s}^0}{g_{\ast s}^\text{d}}{\frac{1}{3}} \, ,
\end{equation}
where we used that $\int c y^{c+n-1} e^{-y^c} dy =1 $ for $c=2$ in both cases, $n=0$ and $n=2$. We obtain a very simple, exact expression.
With~\eqref{rmsnumomentum} this implies
\begin{eqnarray}
\gamma_\text{ld}^\text{rms} &=&
\fb{\zeta (3)}{15 \zeta (5)}{\frac{1}{2}}
\fb{11}{4}{\frac{1}{3}}
\fb{g_{\ast s}^0}{g_{\ast s}^\text{d}}{\frac{1}{3}}
\frac{p_\text{ini}}{T_\text{d}} \nonumber \\
&\simeq& 0.389 \, \frac{p_\text{ini}}{T_\text{d}} \fb{g_{\ast s}^0}{g_{\ast s}^\text{d}}{\frac{1}{3}} ,
\end{eqnarray}
which can be expressed as
\begin{equation}
\label{grmsdec}
\gamma_\text{ld}^\text{rms} \simeq 0.0105 \, \frac{\DN_\text{eff}}{\Omega_\text{ld}^0 h^2} \frac{g_\text{hdm}}{2}
\end{equation}
by lower moments. Again, this must be, since there are no further parameters. All higher momenta are fixed by the first two.
\begin{figure}
\includegraphics[width=\columnwidth]{f_p__overplot.pdf}
\caption{Momentum distribution functions with perfect mimicry in the first two cosmological observables corresponding to their first two moments and their next higher moment.
Depicted are distributions for the early decay (black, solid), the late decay (black, dashed), Pauli-Dirac (black, dash-dotted), Einstein-Bose (grey, dash-dotted) and decay in a matter-dominated universe (grey, dashed). Zeroth and first moment have been set to unity. Vertical lines mark the corresponding second moments (= skewness). Units are arbitrary.
}
\label{fig:fp_overplot}
\end{figure}
The predicted deviation between \ensuremath{\nu_s} HDM and dpHDM (with $g_\text{hdm}=2$) in the normalised root-mean-squares of absolute momenta
is thus found by combining~\eqref{grmsdec} and~\eqref{grmsnus} as
\begin{equation}
\gamma_\text{ld}^\text{rms}/\gamma_{\ensuremath{\nu_s}}^\text{rms} \simeq 0.0105 \, \frac{\DN_\text{eff}^\frac{3}{4}}{\Omega_\text{dec}^0 h^2} \simeq 0.989 \, \frac{\DN_\text{eff}^\frac{3}{4}}{m_\text{hdm}^\text{eff}/\text{eV}} \, .
\end{equation}
Under the crude assumption of Gaussian error propagation this implies for~\eqref{hdmsignal} a prediction,
\begin{equation}
\gamma_\text{ld}^\text{rms}/\gamma_{\ensuremath{\nu_s}}^\text{rms} = 1.66 \pm 0.81 \; (\text{1-}\sigma)\, .
\end{equation}
Since the momentum distribution functions differ between early and late decay as can be seen in Fig.~\ref{fig:fp_overplot}, these numbers differ between these two cases. We find $ \langle p_\text{ed} \rangle_\text{rms} \simeq 1.03 \times \langle p_\text{ld} \rangle_\text{rms}$ and thus
\begin{equation}
\gamma_\text{ed}^\text{rms}/\gamma_{\ensuremath{\nu_s}}^\text{rms} = 1.72 \pm 0.84 \; (\text{1-}\sigma)\, .
\end{equation}
So there is a 66\% (72\%) difference in the mean for the late (early) decay but the uncertainty of the prediction is decisively larger.
Obviously, the difference to \ensuremath{\nu_s} HDM is too small to be detected in an extended analysis that would have to try to determine $\gamma^\text{rms}$ as additional free parameter.
Not to mention that the observations need to be sensitive to this effect.
We do not expect that current observations or observations in the near future will be able to distinguish dpHDM from \ensuremath{\nu_s} HDM in all cases. In this sense, the mimicry can be ''perfect for our limited observational capabilities.``\footnote{
This usage of the notion of mimicry is actually closest to mimicry in biology.
}
It is tempting to look for a first hint if this negative conclusion could change in the next decade. With full Planck data available the galaxy survey EUCLID (to be launched 2020) will increase sensitivities dramatically by roughly an order of magnitude, cf.~\cite{Hamann:2012fe}. Including galaxy surveys, eBOSS and DESI, and a new (Stage-IV) CMB polarimeter,
expected sensitivities go down to $\sigma(\sum m_\nu) = 16 \text{ meV}$ and $\sigma(N_\text{eff})= 0.020$~\cite{Abazajian:2013oma}.
These sensitivities would imply 1-$\sigma$ errors in the prediction of $\gamma_\text{dec}^\text{rms}/\gamma_{\ensuremath{\nu_s}}^\text{rms}$ that are nine times smaller than the predicted deviation for the current mean values~\eqref{hdmsignal}. However, these means can and will change, but for means half the size the 1-$\sigma$ errors were still $3.5$ times smaller than the predicted deviation.
If not from testing our standard interpretation alone, the opportunity to distinguish $\ensuremath{\nu_s}$HDM from dpHDM, motivates to study which future observations can how far probe the root-mean-square of absolute momenta of a free-streaming population.
\section{Summary and Conclusions}
\label{sec:conclusions}
\begin{table}
\begin{tabular}{l|c|c|c}
\shortstack{daughter \\ masses} & \shortstack{mimicry in \\ 1st two} & opportunity? & \ensuremath{\nu_s}\ equivalent \\ \hline
$m_1 \ll m_2$ & \checkmark & $2\times$larger $T^\text{nr}$ & \shortstack{2\ensuremath{\nu_s} : $T_1=T_2$, \\ $m_2 >m_1=0$} \\ \hline
$m_1=m_2$ & \checkmark & \shortstack{breaks down \\ in $\langle p \rangle_\text{rms}$} & \shortstack{''perfect`` for \\ current obs.} \\ \hline
$m_1\lesssim m_2$ & \checkmark & as \ensuremath{\nu_s}\ equivalent & \shortstack{2\ensuremath{\nu_s} : $T_1=T_2$, \\ $m_2 \gtrsim m_1>0$} \\
\end{tabular}
\caption{Overview of different cases depending on the daughter masses. In every case there is mimicry in the first two cosmological observables. Given are opportunities to distinguish these cases from \ensuremath{\nu_s} HDM and how they are represented by thermal \ensuremath{\nu_s}\ populations.}
\label{tab:summary}
\end{table}
As summarised in Tab.~\ref{tab:summary}, it depends on the daughter masses where and how the mimicry breaks down. Every case can be implemented easily as \ensuremath{\nu_s}\ populations in existing numerical tools.
If the daughters do posses a large mass hierarchy, the temperature when dpHDM becomes non-relativistic is larger by a factor of two compared to \ensuremath{\nu_s} HDM, so that this case might be in stronger tension with CMB data.
If the daughters are mass degenerate, they are indistinguishable from \ensuremath{\nu_s} HDM in analyses like~\cite{Hamann:2013iba}.
Connecting cosmological ''observables`` with moments of the momentum distribution functions depicted in Fig.~\ref{fig:fp_overplot}, we find that for mass-degenerate daughters the mimicry breaks down only, if the next higher moment of the momentum distribution, the skewness, is considered. We define the \textit{normalised root-mean-square of absolute momenta} and find sizeable differences in its predicted value between dpHDM and \ensuremath{\nu_s} HDM. While these are certainly too small for current observations, this is a \textit{qualitative difference} compared to the attempt to distinguish different \ensuremath{\nu_s} HDM models, where the mimicry is perfect to arbitrary order of moments.
Other opportunities depend on the time of decay: For certain times during the BBN era, dpHDM becomes non-relativistic later than \ensuremath{\nu_s} HDM. A decay after BBN increases $N_\text{eff}|_\text{cmb} > N_\text{eff}|_\text{bbn}= N_\text{eff}^\text{sm}$, which motivates analyses that drop the BBN consistency relation.
To conclude, current cosmological observations are sensitive to sub-eV, not fully-thermalised HDM characterised by $\DN_\text{eff} < 1$ and $m_\text{hdm}^\text{eff} < \text{ eV}$.
In that case, from a cosmological point of view \ensuremath{\nu_s} HDM is not preferred over dpHDM, neither from theoretical nor practical simplicity.
Our (current) blindness for the case of mass-degenerate daughters, should prevent us from premature conclusions when interpreting signals like~\eqref{hdmsignal}. Fortunately, there are various cases that can be considered easily and gainfully in likelihood analyses utilising available data already.
After our proof of principle that, in contrast to different, thermal \ensuremath{\nu_s} HDM models, the mimicry of dpHDM breaks down at least in the root-mean-square of absolute momenta, it is an open question which observation due to which effect will be able to distinguish the two possibilities.
\subsection*{Acknowledgements}
\noindent
I would like to thank Jan Hamann for valuable discussions.
I acknowledge support from the German Academy of Science through the Leopoldina Fellowship Programme
grant LPDS 2012-14.
\begin{appendix}
\section{Free-streaming scale and transition time}
\label{appendix:A}
Starting from (93) in~\cite{Lesgourgues:2006nd},
\begin{equation}
k^\text{fs}(T) =\left(\frac{3}{2} \frac{H^2(T) a^2(T)}{\langle v \rangle(T)}\right)^2 \text{, } \lambda^\text{fs}(T)=2\pi\frac{a(T)}{k^\text{fs}(T)},
\end{equation}
we approximate the average absolute velocity of the population $\langle v \rangle = \langle p\rangle/m$ at the transition, $\langle p \rangle =m$, as the speed of light $c$. We assume that it then decreases as $a^{-1} \propto T$, so that $\langle v \rangle(T) = c \, T/T^\text{nr} $, if $T^\text{nr}$ denotes the temperature of the Universe at the transition.\footnote{
As a side note,
we could also calculate $\langle v \rangle =\langle E_\text{kin} \rangle /m = \rho_\text{kin}/ (mn)$
and insert appropriate expressions for the kinetic energy density $\rho_\text{kin}$.
}
The scale factor can be expressed in temperatures as $a(T)/a_0=T_0/T$ and $H^2 = H_0^2 \left(\Omega_\text{m} (T/T_0)^3 +\Omega_\Lambda\right)$.
Inserting yields
\begin{equation}
k^\text{fs}(T) \simeq 4.08 \times 10^{-4} \left(\Omega_\text{m} \fb{T}{T_0}{3} +\Omega_\Lambda \right)^\frac{1}{2} \frac{T_0 T^\text{nr}}{T^2} \frac{h}{\text{Mpc}}
\end{equation}
and
\begin{equation}
\lambda^\text{fs}(T) \simeq 1.54 \times 10^4 \left(\Omega_\text{m} \fb{T}{T_0}{3} +\Omega_\Lambda \right)^{-\frac{1}{2}} \frac{T}{T^\text{nr}} \frac{\text{Mpc}}{h} .
\end{equation}
For $k^\text{nr}=k^\text{fs}(T^\text{nr})$ we obtain~\eqref{knr}.
For reference we provide the often used free-streaming scale of the population today,
\begin{equation}
\lambda^\text{fs} \simeq 1.54 \times 10^4 \frac{T_0}{T^\text{nr}} h^{-1} \text{ Mpc}.
\end{equation}
Consistently we see from (93) of~\cite{Lesgourgues:2006nd}
\begin{equation}
\lambda^\text{fs}(t^\text{nr}) =2 \pi (2/3)^{1/2} \langle v(t^\text{nr}) \rangle/H(t^\text{nr})=\sqrt{6} \pi t^\text{nr} \, ,
\end{equation}
where we used $\langle v(t^\text{nr}) \rangle =1$ and $H=2/(3 t)$ in matter domination. We calculate the time when a population becomes non-relativistic in a Universe filled
with radiation and matter as
$
t^\text{nr} = \frac{{t_\text{eq}}}{2-\sqrt{2}} \left( \left(\frac{T_\text{eq}}{T^\text{nr}} -2 \right) \left( \frac{T_\text{eq}}{T^\text{nr}} +1 \right)^{1/2} +2 \right)
$,
where $T_\text{eq}$ $({t_\text{eq}})$ denotes the temperature (time) at matter-radiation equality.
\section{Neutrino distribution specified by its moments}
\label{appendix:nudistr}
This appendix shall improve the accessibility of Sec.~\ref{sec:breakdown} for the reader.
We repeat findings of Cuoco, Lesgourgues, Mangano and Pastor in~\cite{Cuoco:2005qr} and refer the reader to
the original work.
The solution to the collisionless kinetic equations in a Lemaitre-Friedman-Robertson-Walker
universe is a Fermi-Dirac function $f(\vec{p}) = 1/(e^{(E-\mu)/T}+1)$ with particle energy $E^2=|\vec{p}|^2 +m^2= p^2+m^2$.
We are interested in the standard situation with
vanishing chemical potential $\mu$ and early times, when neutrinos are ultra-relativistic.
Considering a radiation dominated universe, where $T\propto a^{-1}$, and defining the comoving momentum $y=pa$
one finds
\begin{equation}
\label{nudistr}
df(p,T_\nu)= \frac{1}{\pi^2} T_\nu^3 \frac{y^2}{e^y+1} dy \, ,
\end{equation}
where isotropy has been exploited to reduce the dimension of the differential and
which is normalised such that the integral yields the number density as required.
One can define a set of moments
\begin{equation}
\label{numoments}
Q^{(n)}_\nu= \frac{1}{\pi^2} \fb{4}{11}{\frac{3+n}{3}} T^{3+n} \int y^{2+n} f(y) dy
\end{equation}
in terms of the neutrino temperature $T_\nu= (4/11)^{1/3} T$. These moments can
specify the neutrino distribution $df(y)$ regardless of the specific case at hand.
In~\cite{Cuoco:2005qr} this is used to explore observation prospects for non-thermal contributions to the
standard neutrino spectrum.
If the distribution decays at large comoving
momentum as $e^{-y}$, it admits moments of all orders.
In~\cite{Cuoco:2005qr} the neutrino distribution is given in terms of its moments as
\begin{equation}
df(y)= \frac{y^2}{e^y+1} \sum_{m=0}^\infty \sum_{k=0}^m
c_k^{(m)} Q^{(k)}_\nu T_\nu^{-k} P_m(y) dy
\end{equation}
with $P_m(y)= \sum_{k=0}^m c_k^{(m)} y^k$, $m$ being the degree of $P_m(y)$ and $c_k^{(m)}$ being a coefficient,
denoting the set of polynomials orthonormal with respect to the measure $y^2/(e^y+1)$, i.e.,
$\int_0^\infty dy \frac{y^2}{e^y+1} P_n(y)P_m(y) =\delta_{nm}$.
For a Fermi-Dirac distribution all moments can be expressed in terms of the lowest moment $Q^{(0)}= n_\nu$ or,
equivalently, as functions of the temperature $T_\nu$ since it is the only independent parameter.
For neutrinos the (first two) cosmological observables can be written as
\begin{align}
N_\text{eff} = \frac{120}{7 \pi^2} T_\nu^{-4} \sum_\alpha Q_{\nu \alpha}^{(1)}
\end{align}
and
\begin{align}
\Omega_\nu^0 &= \frac{\sum m_\nu}{\rho_{\text{c}}} Q_\nu^{(0)} \fb{T_0}{T}{3} \nonumber \\
&\Leftrightarrow \Omega_\nu^0 h^2 \simeq 0.162 \, \frac{\sum m_\nu}{\text{eV}} T^{-3} Q_\nu^{(0)} ,
\end{align}
where it is assumed that the three neutrinos share the same distribution today and
the small correction from $e^+e^-$ annihilation, $N_\text{eff}^\text{sm}=3.046 \neq 3$, is incorporated in the numerical prefactor.
In~\cite{Cuoco:2005qr} $N_\text{eff}$ and $f_\nu$ are used to probe deviations from the standard neutrino spectrum
in the first two moments.
Also the following facts are used in our discussion:
The average absolute neutrino momentum $\langle p_\nu \rangle$ expressed by moments of their distribution~\eqref{numoments} reads
\begin{equation}
\label{averagenumomentum}
\langle p_\nu \rangle = \frac{Q_\nu^{(1)}}{Q_\nu^{(0)}}
= \frac{7 \pi^4}{180 \zeta(3)} \fb{4}{11}{\frac{1}{3}} T \simeq 3.15 \, T_\nu
\end{equation}
and the root-mean-square of absolute momenta $\langle p_\nu \rangle_\text{rms} =(Q_\nu^{(2)}/Q_\nu^{(0)})^{1/2}$ can be calculated analogously as
\begin{equation}
\label{rmsnumomentum}
\langle p_\nu \rangle_\text{rms}
= \fb{15 \zeta(5)}{\zeta(3)}{\frac{1}{2}} \fb{4}{11}{\frac{1}{3}} T \simeq 3.6 \, T_\nu \, .
\end{equation}
\end{appendix}
\phantomsection
\addcontentsline{toc}{chapter}{References}
|
\section{Introduction}
\label{intro}
Since the pioneering studies \cite{klein_prc,gluon,strikman} on diffractive vector meson production in ultra peripheral heavy ion collisions (UPHIC) about fourteen years ago, a large number of papers on the subject has been published considering several improvements in the theoretical description \cite{outros_klein,vicmag_mesons1,outros_vicmag_mesons,outros_frankfurt,Schafer,vicmag_update,gluon2,motyka_watt,Lappi,griep,Guzey,Martin,cisek} and experimental analysis \cite{cdf,star,phenix,alice,alice2,lhcb}. Recent experimental results from CDF \cite{cdf} at Tevatron, STAR \cite{star} and PHENIX \cite{phenix} at RHIC and ALICE \cite{alice,alice2} and LHCb \cite{lhcb,lhcb2} at LHC have demonstrated that the study of photon-hadron interactions in these colliders is feasible and that the data can be used to constrain the description of the hadronic structure at high energies (For reviews see Ref. \cite{upc}). In particular, the recent theoretical studies performed in Refs. \cite{Guzey,Martin} have demonstrated that the description of the ALICE and LHCb data using the collinear factorization formalism is strongly dependent on the choice of the parameterization of the gluon distribution and on the magnitude of the nuclear shadowing effects (For a similar analysis using the $k_T$-factorization approach see Ref. \cite{cisek}). On the other hand, as originally proposed in Ref. \cite{vicmag_mesons1}, the diffractive photoproduction of vector mesons in hadronic collisions can also be studied in the color dipole formalism \cite{nik} and used to constrain the magnitude of the nonlinear effects which are predicted to be present at high parton densities by the QCD dynamics at high energies. In this formalism the predictions for the vector meson production in hadronic collisions are strongly dependent on the model used to calculate the forward dipole - target scattering amplitude $\cal{N}$. As the current color dipole predictions for the diffractive photoproduction of $J/\Psi$ in hadronic collisions \cite{outros_vicmag_mesons,vicmag_update,motyka_watt,Lappi,griep} consider distinct models for this quantity, photon spectrum, skewness corrections and/or for the vector meson wavefunction, a direct comparison between its predictions is a hard task. Our goal in this paper is to compare the predictions obtained considering different models for $\cal{N}$ assuming a unified treatment for the photon spectrum and $J/\Psi$ wavefunction.
It allow us to estimate the theoretical uncertainty present in the current predictions in the literature. After showing that the different models used in our calculations are able to describe the $\gamma p$ HERA data, we present predictions for the photonuclear $J/\Psi$ production which could be probed in future electron - ion colliders \cite{eA}. Moreover, we compare our predictions for the diffractive
photoproduction of $J/\Psi$ in $pp$ and $PbPb$ collisions with recent LHCb and ALICE data, respectively. Finally, we present predictions for future LHC energies and $pPb$ collisions.
The paper is organized as follows. In the next section we present a brief review of photon - hadron
interactions in $pp$ and $PbPb$ collisions, as well as of the diffractive photoproduction of $J/\Psi$ in the color dipole formalism. We also present the models for the dipole - target scattering
amplitude used in our calculations. In Section \ref{resultados} we present our predictions
for the diffractive photoproduction of $J/\Psi$ in $pp/pPb/PbPb$ collisions and a comparison
with the ALICE and LHCb data is also shown. Finally, in Section \ref{sumario} we summarize our main conclusions.
\section{Formalism}
\label{formulas}
Initially let us present a brief review of photon - hadron interactions in hadronic collisions.
Let us consider a hadron-hadron interaction at large impact parameter
($b > R_{h_1} + R_{h_2}$) and at ultra relativistic energies. In this regime we expect the
electromagnetic interaction to be dominant.
In heavy ion colliders, the heavy nuclei give rise to strong electromagnetic fields due to
the coherent action of all protons in the nucleus. In a similar way, this kind of interaction
also occurs in ultra relativistic protons in $pp(\bar{p})$ colliders.
The photon stemming from the electromagnetic field
of one of the two colliding hadrons can interact directly with the other hadron
(photon-hadron process). The total
cross section for a given process can be factorized in terms of the equivalent flux of
photons of the hadron projectile and the photon-target production cross
section \cite{upc}. The cross section for diffractive $J/\Psi$ photoproduction in a
hadron-hadron collision is given by,
\begin{eqnarray}
\sigma (h_1 h_2 \rightarrow h_1 \otimes J/\Psi \otimes h_2) =
\int \limits_{\omega_{min}}^{\infty} d\omega \,
\frac{dN_{\gamma/h_1}(\omega)}{d\omega}\,
\sigma_{\gamma h_2 \rightarrow J/\Psi h_2} (W_{\gamma h_2}^2) \nonumber \\ +
\int \limits_{\omega_{min}}^{\infty} d\omega \,
\frac{dN_{\gamma/h_2}(\omega)}{d\omega}\,
\sigma_{\gamma h_1 \rightarrow J/\Psi h_1} (W_{\gamma h_1}^2),
\label{sighh}
\end{eqnarray}
where $\otimes$ represents a rapidity gap in the final state, $\omega$ is the photon
energy in the collider frame, $\omega_{min}=M_{J/\Psi}^2/4\gamma_L m_p$, $\gamma_L$ is the Lorentz boost of a
single beam, $\frac{dN_{\gamma}}{d\omega}$ is the equivalent photon flux,
$W_{\gamma h}^2=2\,\omega \sqrt{s_{\mathrm{NN}}}$ and $\sqrt{s_{\mathrm{NN}}}$ is
the c.m.s energy of the
hadron-hadron system.
Considering the requirement that photoproduction
is not accompanied by hadronic interaction (ultra-peripheral
collision) an analytic approximation for the equivalent photon flux of a nuclei can be
calculated and given by \cite{upc}
\begin{eqnarray}
\frac{dN_{\gamma/A}\,(\omega)}{d\omega}= \frac{2\,Z^2\alpha_{em}}{\pi\,\omega}\, \left[\bar{\eta}\,K_0\,(\bar{\eta})\, K_1\,(\bar{\eta})+ \frac{\bar{\eta}^2}{2}\,{\cal{U}}(\bar{\eta}) \right]\,
\label{fluxint}
\end{eqnarray}
where $\bar{\eta}=\omega\,(R_{h_1} + R_{h_2})/\gamma_L$ and
${\cal{U}}(\bar{\eta}) = K_1^2\,(\bar{\eta})- K_0^2\,(\bar{\eta})$.
Eq. (\ref{fluxint}) will be used in our calculations of the diffractive photoproduction of $J/\Psi$ in $pPb$ and $PbPb$
collisions. On the other hand, for proton-proton collisions, we assume that the photon
spectrum of a relativistic proton is given by \cite{Dress},
\begin{eqnarray}
\frac{dN_{\gamma/p}(\omega)}{d\omega} = \frac{\alpha_{\mathrm{em}}}{2 \pi\, \omega} \left[ 1 + \left(1 -
\frac{2\,\omega}{\sqrt{s_{NN}}}\right)^2 \right]
\left( \ln{\Omega} - \frac{11}{6} + \frac{3}{\Omega} - \frac{3}{2 \,\Omega^2} + \frac{1}{3 \,\Omega^3} \right) \,,
\label{eq:photon_spectrum}
\end{eqnarray}
with the notation $\Omega = 1 + [\,(0.71 \,\mathrm{GeV}^2)/Q_{\mathrm{min}}^2\,]$ and
$Q_{\mathrm{min}}^2= \omega^2/[\,\gamma_L^2 \,(1-2\,\omega /\sqrt{s_{NN}})\,] \approx
(\omega/
\gamma_L)^2$.
The main input in our calculations is the diffractive photoproduction cross section $\sigma_{\gamma h \rightarrow J/\Psi h}$.
In what follows we describe the $\gamma h$ scattering in the dipole frame, in which most of the energy
is
carried by the hadron, while the photon has
just enough energy to dissociate into a quark-antiquark pair
before the scattering. In this representation the probing
projectile fluctuates into a
quark-antiquark pair (a dipole) with transverse separation
$\mbox{\boldmath $r$}$ long before the interaction, which then
scatters off the hadron \cite{nik}. Initially, let us consider a photon - proton interaction ($h = p$).
In the dipole picture the amplitude for the diffractive photoproduction of an exclusive final state,
such as a $J/\Psi$, in a $\gamma p$ collision is given by
(See e.g. Refs. \cite{nik,vicmag_mesons,KMW,armesto_amir})
\begin{eqnarray}
{\cal A}^{\gamma p \rightarrow J/\Psi p}(x,\Delta) = i
\int dz \, d^2\mbox{\boldmath $r$} \, d^2\mbox{\boldmath $b$} e^{-i[\mbox{\boldmath $b$}-(1-z)\mbox{\boldmath $r$}].\mbox{\boldmath $\Delta$}}
\,\, (\Psi_{J/\Psi}^* \Psi) \,\,2 {\cal{N}}_p(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$})
\label{sigmatot2}
\end{eqnarray}
where $(\Psi_{J/\Psi}^* \Psi)$ denotes the overlap of the photon and $J/\Psi$
transverse wave functions. The variable $z$ $(1-z)$ is the
longitudinal momentum fractions of the quark (antiquark), $\Delta$ denotes the transverse
momentum lost by the outgoing proton ($t = - \Delta^2$) and $x$ is the Bjorken variable.
The variable $\mbox{\boldmath $b$}$ is the transverse distance from the center of the target to the center of mass of the
$q \bar{q}$ dipole and the factor in the exponential arises when one takes into account
non-forward corrections to the wave functions \cite{non}.
Moreover, ${\cal{N}}_p (x,\mbox{\boldmath $r$},\mbox{\boldmath $b$})$ denotes the forward amplitude of the scattering of a dipole of size $\mbox{\boldmath $r$}$ on the proton, which is directly related to the QCD
dynamics (see below). The total cross section for diffractive photoproduction of $J/\Psi$ is given by
\begin{eqnarray}
\sigma (\gamma p \rightarrow J/\Psi p) = \frac{1}{16\pi} \int dt \, |{\cal{A}}^{\gamma p \rightarrow J/\Psi p}(x,\Delta)|^2\,R_g^2(1 + \beta^2)\,\,
\label{totalcs}
\end{eqnarray}
where $\beta$ is the ratio of real to imaginary parts of the scattering
amplitude, which can be obtained using dispersion relations
$Re {\cal A}/Im {\cal A}=\mathrm{tan}\,(\pi \lambda_e/2)$. Moreover, $R_g$ is the skewness factor, which is associated to the fact that the gluons attached to the $q\bar{q}$ pair can carry different light-cone fractions $x$, $x^{\prime}$ of the proton. In the limit that $x^{\prime} \ll x \ll 1$ and at small $t$ and assuming that the gluon density has a power-law form, it is given by \cite{Shuvaev:1999ce}
\begin{eqnarray}
\label{eq:Rg}
R_g(\lambda_e) = \frac{2^{2\lambda_e+3}}{\sqrt{\pi}}
\frac{\Gamma(\lambda_e+5/2)}{\Gamma(\lambda_e+4)},
\quad\text{with} \quad \lambda_e \equiv
\frac{\partial\ln\left[\mathcal{A}(x,\,\Delta)\right]}{\partial\ln(1/x)}.
\end{eqnarray}
As demonstrated in Ref. \cite{harland} the estimate obtained using this approximation for $R_g$ is strongly dependent on the parton distribution used in the calculation. Furthermore, the incorporation of the skewness correction at small-$x$ in the dipole models still is an open question. Consequently, the factor $R_g$ as given in Eq. (\ref{eq:Rg}) should be considered a phenomenological estimate.
If we assume that the nucleus
scatters elastically (coherent production) and also that the scattering happens
in the high energy regime (large coherence length: $l_c \gg R_A$) the total photon-nucleus cross
section is given by:
\cite{vmprc,kop1}
\begin{eqnarray}
\sigma^{coh}\, (\gamma A \rightarrow J/\Psi A) = \int d^2\mbox{\boldmath $b$} \left[
\int d^2\mbox{\boldmath $r$}
\int dz (\Psi_{J/\Psi}^* \Psi) \, \mathcal{N}_A(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$})\right]^2
\label{totalcscoe}
\end{eqnarray}
where $ {\cal N}_A (x, \mbox{\boldmath $r$}, \mbox{\boldmath $b$})$ is the forward dipole-nucleus scattering amplitude.
Finally, the photon wave functions appearing in Eq. (\ref{sigmatot2}) are well known in literature
\cite{KMW}. For the meson wave function, we have considered the Gauss-LC model \cite{KMW}
which is a simplification of the DGKP wave functions \cite{dgkp}. The motivation for this choice is its
simplicity and the fact that, for the present purposes, the final results are not very sensitive
these details. In photoproduction, this leads only to an uncertainty of a few percent in the
overall normalization. The parameters of the meson wave function can be found in Ref. \cite{KMW}.
The scattering amplitude ${\cal{N}}(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$})$ contains all
information about the target and the strong interaction physics.
In the Color Glass Condensate (CGC) formalism \cite{CGC,BAL}, it encodes all the
information about the
non-linear and quantum effects in the hadron wave function. It can be obtained by solving
an appropriate evolution
equation in the rapidity $Y\equiv \ln (1/x)$, which in its simplest form is the
Balitsky-Kovchegov (BK) equation \cite{BAL,kov}. In recent years, the running coupling corrections to BK evolution kernel were explicitly calculated \cite{kovwei1,balnlo}, including the $\alpha_sN_f$ corrections to
the kernel to all orders, and its solution studied in detail \cite{javier_kov,javier_prl}. Basically, one has that the running of the coupling reduces the speed of the evolution to values compatible with experimental $ep$ HERA data \cite{bkrunning,weigert}. In Ref. \cite{bkrunning} the translational invariance approximation was assumed, which implies ${\cal{N}}(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$}) = {\cal{N}}(x,\mbox{\boldmath $r$}) S(\mbox{\boldmath $b$})$, with the normalization of the dipole cross section being fitted to data and two distinct initial conditions, inspired in the Golec Biernat-Wusthoff (GBW) \cite{GBW} and McLerran-Venugopalan (MV) \cite{MV} models, were considered. The predictions resulted to be almost independent of the initial conditions and, besides, it was observed that it is impossible to describe the experimental data using only the linear limit of the BK equation, which is equivalent to the Balitsky-Fadin-Kuraev-Lipatov (BFKL) equation \cite{bfkl}. In what follows we use the solution obtained in Ref. \cite{bkrunning} considering the MV initial condition, denoting the corresponding predictions by rcBK hereafter. It is important to emphasize that although a complete analytical solution of the BK equation is still lacking, its main properties are known: (a) for the interaction of a small dipole ($|\mbox{\boldmath $r$}| \ll
1/Q_s$), ${\cal{N}}(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$}) \approx \mbox{\boldmath $r$}^2$, implying that
this system is weakly interacting; (b) for a large dipole ($|\mbox{\boldmath $r$}| \gg
1/Q_s$), the system is strongly absorbed and therefore
${\cal{N}}(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$}) \approx 1$. The typical momentum scale, $Q_s^2\propto x^{-\lambda}\,(\lambda\approx 0.3)$, is the so called saturation scale. This property is associated to the
large density of saturated gluons in the hadron wave function. In the last years, several groups have constructed phenomenological models which satisfy the asymptotic behavior of the BK equation in order to fit the HERA and RHIC data (See e.g. Refs. \cite{GBW,dipolos10,iim,kkt,dhj,Goncalves:2006yt,buw,KMW}).
For comparison, in what follows we will also use the GBW model \cite{GBW}, which assumes that
${\cal{N}}_p(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$}) = {\cal{N}}_p(x,\mbox{\boldmath $r$}) S(\mbox{\boldmath $b$})$ with ${\cal{N}}_p(x,\mbox{\boldmath $r$})=1-e^{-\mbox{\boldmath $r$}^2Q_{s,p}^2(Y)/4}$ and $Q_{s,p}^2(Y)=\left(x_0/x\right)^{\lambda}$, with the parameters $x_0$ and $\lambda$ determined by the fit to the HERA data. Moreover, we also consider the
b-CGC model proposed in Ref. \cite{KMW}, which improves the Iancu - Itakura - Munier (IIM) model
\cite{iim} with the inclusion of the impact parameter dependence in the dipole - proton scattering amplitude. Following \cite{KMW} we have:
\begin{eqnarray}
\mathcal{N}_p(x,\mbox{\boldmath $r$},{\mbox{\boldmath $b$}}) =
\left\{ \begin{array}{ll}
{\mathcal N}_0\, \left(\frac{ r \, Q_{s,p}}{2}\right)^{2\left(\gamma_s +
\frac{\ln (2/r Q_{s,p})}{\kappa \,\lambda \,Y}\right)} & \mbox{$r Q_{s,p} \le 2$} \\
1 - \exp^{-A\,\ln^2\,(B \, r \, Q_{s,p})} & \mbox{$r Q_{s,p} > 2$}
\end{array} \right.
\label{eq:bcgc}
\end{eqnarray}
with $Y=\ln(1/x)$ and $\kappa = \chi''(\gamma_s)/\chi'(\gamma_s)$, where $\chi$ is the
LO BFKL characteristic function. The coefficients $A$ and $B$
are determined uniquely from the condition that $\mathcal{N}_p(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$})$, and its derivative
with respect to $rQ_s$, are continuous at $rQ_s=2$.
In this model, the proton saturation scale $Q_{s,p}$ depends on the impact parameter:
\begin{equation}
Q_{s,p}\equiv Q_{s,p}(x,{\mbox{\boldmath $b$}})=\left(\frac{x_0}{x}\right)^{\frac{\lambda}{2}}\;
\left[\exp\left(-\frac{{b}^2}{2B_{\rm CGC}}\right)\right]^{\frac{1}{2\gamma_s}}.
\label{newqs}
\end{equation}
The parameter $B_{\rm CGC}$ was adjusted to give a good
description of the $t$-dependence of exclusive $J/\psi$ photoproduction.
The factors $\mathcal{N}_0$ and $\gamma_s$ were taken to be free. In this
way a very good description of $F_2$ data was obtained.
One of the parameter set which is going to be used here is the one presented in the second line of
Table II of \cite{watt}: $\gamma_s = 0.46$, $B_{CGC} = 7.5$ GeV$^{-2}$,
$\mathcal{N}_0 = 0.558$, $x_0 = 1.84 \times 10^{-6}$ and $\lambda = 0.119$.
More recently, the parameters of this model have been updated in Ref. \cite{amir} (considering the
recently released high precision combined HERA data), becoming $\gamma_s = 0.6599$, $B_{CGC} = 5.5$ GeV$^{-2}$,
$\mathcal{N}_0 = 0.3358$, $x_0 = 0.00105 \times 10^{-5}$ and $\lambda = 0.2063$.
In what follows we will use these two sets of parameters in our calculations, with the resulting predictions being denoted bCGC and bCGC NEW, respectively.
In order to estimate the diffractive photoproduction of $J/\Psi$ in $PbPb$ collisions we need to specify the
forward dipole - nucleus scattering amplitude, $\mathcal{N}_A(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$})$.
Following \cite{vmprc} we will use in our calculations the model proposed in Ref.
\cite{armesto}, which describes the current experimental data on the nuclear
structure function as well as includes the impact parameter dependence in the dipole
nucleus cross section. In this model the forward dipole-nucleus amplitude is given by
\begin{eqnarray}
{\cal{N}}_A(x,\mbox{\boldmath $r$},\mbox{\boldmath $b$}) = 1 - \exp \left[-\frac{1}{2} \, \sigma_{dp}(x,\mbox{\boldmath $r$}^2)
\,T_A(\mbox{\boldmath $b$})\right] \,\,,
\label{enenuc}
\end{eqnarray}
where $\sigma_{dp}$ is the dipole-proton cross section given by
\begin{eqnarray}
\sigma_{dp} (x,\mbox{\boldmath $r$}^2) = 2 \int d^2\mbox{\boldmath $b$} \,\,\mathcal{N}_p(x,\mbox{\boldmath $r$},{\mbox{\boldmath $b$}})
\end{eqnarray}
and $T_A(\mbox{\boldmath $b$})$ is the nuclear profile
function, which is obtained from a 3-parameter Fermi distribution for the nuclear
density normalized to $A$.
The above equation
sums up all the
multiple elastic rescattering diagrams of the $q \overline{q}$ pair
and is justified for large coherence length, where the transverse separation $\mbox{\boldmath $r$}$ of
partons
in the multiparton Fock state of the photon becomes a conserved quantity, {\it i.e.}
the size
of the pair $\mbox{\boldmath $r$}$ becomes eigenvalue
of the scattering matrix. In what follows we will compute $\mathcal{N}_A$ considering
the different models for the dipole - proton scattering amplitude discussed before.
\begin{figure}
\begin{tabular}{cc}
\includegraphics[scale=0.35]{jpsi_gamap_paper.eps} &
\includegraphics[scale=0.35]{jpsi_coh_paper.eps}
\end{tabular}
\caption{(Color online) Diffractive photoproduction of $J/\Psi$ in (a) $\gamma p$
and (b) $\gamma Pb$ collisions. Data from HERA \cite{hera_data}. }
\label{fig1}
\end{figure}
\section{Results}
\label{resultados}
In Fig. \ref{fig1} we present our predictions for the diffractive photoproduction of $J/\Psi$ in (a) $\gamma p$
and (b) $\gamma Pb$ collisions considering as input the different models for the dipole - proton scattering amplitude discussed in the previous section. For the $\gamma p$ case , we
find that the distinct models are able to describe the low energy data, but predict different behaviours at high energies. It is important to emphasize that the parameterization for the solution of the BK equation proposed in Ref. \cite{bkrunning} is only valid for $x \le 10^{-2}$. This restricts its use for $W_{\gamma p} \ge 30$ GeV. Moreover, we find that the improved version of the bCGC (bCGC NEW) predicts a larger growth for the total cross section in comparison with the previous one, which becomes its predictions compatible with the HERA data. For $\gamma Pb$ collisions, we find that the rcBK and bCGC predictions are similar at low energies, but differ by a factor 1.8 at larger energies. On the other hand, the bCGC NEW prediction can be considered a lower bound for the diffractive photoproduction of $J/\Psi$ at low energies, while the GBW one is an upper bound. As already pointed in Ref. \cite{vmprc}, future measurements in an $eA$ collider \cite{eA} could discriminate between these models.
Let us now estimate the rapidity distribution in diffractive photoproduction
of $J/\Psi$ in hadronic collisions.
The rapidity ($Y$) distribution of the $J/\Psi$ in the final state can be directly computed from Eq. (\ref{sighh}), by using its relation with the photon energy $\omega$, i.e. $Y\propto \ln \, ( \omega/m_{J/\Psi})$. Explicitly, the rapidity distribution is written down as,
\begin{eqnarray}
\frac{d\sigma \,\left[h_1 + h_2 \rightarrow h_1 \otimes J/\Psi \otimes h_2\right]}{dY} = \left[\omega \frac{dN}{d\omega}|_{h_1}\,\sigma_{\gamma h_2 \rightarrow J/\Psi \otimes h_2}\left(\omega \right)\right]_{\omega_L} + \left[\omega \frac{dN}{d\omega}|_{h_2}\,\sigma_{\gamma h_1 \rightarrow J/\Psi \otimes h_1}\left(\omega \right)\right]_{\omega_R}\,
\label{dsigdy}
\end{eqnarray}
where $\otimes$ represents the presence of a rapidity gap in the final state and $\omega_L \, (\propto e^{-Y})$ and $\omega_R \, (\propto e^{Y})$ denote photons from the $h_1$ and $h_2$ hadrons, respectively. As the photon fluxes, Eqs. (\ref{fluxint}) and (\ref{eq:photon_spectrum}), have support at small values of $\omega$, decreasing exponentially at large $\omega$, the first term on the right-hand side of the Eq. (\ref{dsigdy}) peaks at positive rapidities while the second term peaks at negative rapidities. Consequently, given the photon flux, the study of the rapidity distribution can be used to constrain the photoproduction cross section at a given energy. Moreover, in contrast to the total rapidity distributions for $pp$ and $PbPb$ collisions, which will be symmetric about midrapidity ($Y=0$), $d\sigma/dY$ will be asymmetric in $pPb$ collisions due to the differences between the fluxes and process cross sections.
It is important to emphasize that in our calculations we will disregard soft interactions which lead to an extra production of particles that destroy the rapidity gap in the final state.
The inclusion of these additional absorption effects can be parametrized in terms of a multiplicative factor denoted rapidity gap survival probability, $S^2$, which corresponds to the probability of the scattered proton not to dissociate due to the secondary interactions.
In Ref. \cite{Schafer} the authors have estimated $S^2$ and obtained that in $pp/p\bar{p}$ collisions it is $ \sim 0.8 - 0.9$, depending on the rapidity of the vector meson (See also Refs. \cite{Guzey,Martin}). Moreover, in the particular case of nuclear interactions, we also disregard a possible reduction of the nuclear photon flux associated to the diffusion edge of the nucleus.
\begin{figure}
\begin{tabular}{cc}
\includegraphics[scale=0.35]{jpsi_pp_7000.eps} &
\includegraphics[scale=0.35]{jpsi_pp_14000.eps}
\end{tabular}
\caption{(Color online) Rapidity distribution for the diffractive photoproduction of $J/\Psi$ in $pp$ collisions at (a) $\sqrt{s}=7$ TeV and (b) $\sqrt{s}=14$ TeV. Data from LHCb Collaboration \cite{lhcb,lhcb2}.}
\label{fig2}
\end{figure}
In Fig. \ref{fig2} we present our predictions for the rapidity distribution for the diffractive photoproduction of $J/\Psi$ in $pp$ collisions at (a) $\sqrt{s}=7$ TeV and (b) $\sqrt{s}=14$ TeV.
Due to the limitation in the $x$-range of the rcBK solution, we are only able to present its predictions for a restricted rapidity range. We obtain that the differences between the predictions observed in Fig. \ref{fig1} are also presented in the rapidity distribution, with the GBW (bCGC) prediction being an upper (lower) bound for the predictions at $Y = 0$. In particular, the predictions differ by $\approx 30$ \% for central rapidities at $\sqrt{s} = 7$ TeV. For the rapidity range probed by the LHCb Collaboration the difference is larger ($\approx 50$ \%). We obtain that the bCGC and bCGC NEW predictions agree with the data from LHCb Collaboration \cite{lhcb,lhcb2}. As demonstrated in Fig. \ref{fig2} (b), these differences increase with the energy. This motivates future experimental analysis of this process in order to constrain the dipole - proton scattering amplitude and, consequently, the QCD dynamics at high energies.
\begin{figure}
\begin{tabular}{cc}
\includegraphics[scale=0.35]{jpsi_aa_2760.eps} &
\includegraphics[scale=0.35]{jpsi_aa_5500.eps}
\end{tabular}
\caption{(Color online) Rapidity distribution for the diffractive photoproduction of $J/\Psi$ in $PbPb$ collisions at (a) $\sqrt{s}=2.76$ TeV and (b) $\sqrt{s}=5.5$ TeV. Data from ALICE Collaboration \cite{alice,alice2}.}
\label{fig3}
\end{figure}
In Fig. \ref{fig3} we present our predictions for the rapidity distribution for the diffractive photoproduction of $J/\Psi$ in $PbPb$ collisions at (a) $\sqrt{s}=2.76$ TeV and (b) $\sqrt{s}=5.5$ TeV. In this case the cross sections are calculated in terms of the dipole - nucleus scattering amplitude given in Eq. (\ref{enenuc}). Similarly to the $pp$ case, we obtain that the distinct predictions largely differ at central rapidities, which is directly associated to the behavior observed in Fig. \ref{fig1} (b) for $\gamma Pb$ collisions. We obtain that the bCGC NEW prediction is able to describe the current ALICE data \cite{alice,alice2}, in contrast with the other predictions which overestimate the data for $Y = 0$. In particular, the rcBK prediction is not able to describe the data, in agreement with the results obtained in Ref. \cite{Lappi}. As observed in Fig. \ref{fig3} (b) the difference between the predictions is amplified at larger energies.
In Fig. \ref{fig4} we present our predictions for the rapidity distribution for the diffractive photoproduction of $J/\Psi$ in $pPb$ collisions at $\sqrt{s}=5$ TeV. As expected, the rapidity distribution is asymmetric about midrapidity ($Y=0$), being dominated by $\gamma p$ interactions, due to the $Z^2$ enhancement present in the nuclear photon spectrum. We observe that the predictions differ by $\approx 35$ \% at $Y= 0$. Finally, in Table \ref{tab} we present our predictions for the total cross section for the diffractive photoproduction of $J/\Psi$ in $pp$, $pPb$ and $PbPb$ collisions at LHC energies. As expected from our analysis of the rapidity distributions, the predictions for the total cross sections are largely distinct.
\section{Summary}
\label{sumario}
The recent experimental data from RHIC, Tevatron and LHC have demonstrated that the study of photon - hadron interactions in hadron - hadron collisions in order to constrain the QCD dynamics at high energies is feasible. They have motivated the proposition of new observables which can be studied in these processes and the improvement its theoretical description. In particular, in the last year, several studies have been performed considering the collinear formalism and the DGLAP evolution, which demonstrated that the diffractive photoproduction of vector mesons can be used to constrain the behaviour of the gluon distribution at small-$x$ and/or the magnitude of the nuclear effects.
However, at the high center-of-mass energies probed in $\gamma h$ interactions at LHC, new dynamics effects associated to nonlinear corrections to the QCD dynamics are expected to be present. These effects are easily included if the process is described in the color dipole formalism, with the cross section being strongly dependent on the model for the dipole - target scattering amplitude. This alternative
approach has been used by several authors in the last years, considering different assumptions for the meson wave function and QCD dynamics, as well as for the free parameters. Our goal in this paper was, using the color dipole formalism, to estimate the theoretical uncertainty associated to description of the QCD dynamics. We have assumed a unique model for the $J/\Psi$ wave function and considered four models for the dipole - proton scattering amplitude. We demonstrated that although these models satisfactorily
describe the HERA data, their predictions are very distinct for the diffractive photoproduction of $J/\Psi$ in $pp/pPb/PbPb$ collisions. In particular, our results point out that the recent ALICE and LHCb data are quite well described by the new version of the bCGC model. Our main conclusion is that the color dipole formalism is able to describe these data and that future measurements can be useful to constrain the magnitude of the nonlinear effects in the QCD dynamics.
\begin{figure}
\includegraphics[scale=0.35]{jpsi_pA_5000.eps}
\caption{(Color online) Rapidity distribution for the diffractive photoproduction of $J/\Psi$ in $pPb$ collisions at $\sqrt{s}=5$ TeV. }
\label{fig4}
\end{figure}
\begin{table}
\begin{center}
\begin{tabular} {||c|c|c|c||}
\hline
\hline
& {\bf GBW}& {\bf bCGC} & {\bf bCGC NEW} \\
\hline
{\bf $pp$} ($\sqrt{s} = 7$ TeV)& 74.0 nb & 49.0 nb & 59.0 nb \\
\hline
{\bf $pp$} ($\sqrt{s} = 14$ TeV)& 113.0 nb & 71.2 nb & 93.7 nb \\
\hline
\hline
{\bf $pPb$} ($\sqrt{s} = 5$ TeV)& 51.3 $\mu$b & 41.0 $\mu$b & 42.8 $\mu$b \\
\hline
\hline
{\bf $PbPb$} ($\sqrt{s} = 2.76$ TeV) & 18.2 mb & 13.6 mb & 11.0 mb \\
\hline
{\bf $PbPb$} ($\sqrt{s} = 5.5$ TeV)& 33.8 mb & 24.4 mb & 20.3 mb \\
\hline
\hline
\end{tabular}
\end{center}
\caption{The total cross section for the diffractive photoproduction of $J/\Psi$ in $pp$, $pPb$ and $PbPb$ collisions at LHC energies.}
\label{tab}
\end{table}
\section*{Acknowledgements}
This work was partially financed by the Brazilian funding agencies CAPES, CNPq, FAPESP and FAPERGS.
|
\section{Introduction}
\label{sec:Intro}
\IEEEPARstart{M}{any} real-world communication systems such as the Internet, World Wide Web (WWW), Peer-to-Peer systems, and Online Social Networks (OSNs) can be modeled as a complex network of interacting dynamical nodes. In the last decade, a considerable amount of research has been done on measuring the characteristics of complex networks in various domains \cite{Newman2006}. To extract useful knowledge from a network, one should study the collected network data. However, the tremendous growth of the Internet and its applications in recent years has resulted in creation of large-scale complex networks involving tens or hundreds of millions of nodes and links.
Thus, it may be impossible or costly to obtain a complete picture of these large networks, and partially-observed data is often being studied for network characterization. The network resulting from such measurements may be thought of as a sample from a larger underlying network. However, sampling in a network context introduces various potential complications such as the bias of selecting the nodes in the procedure of sampling \cite{Frank2011}.
Many network characterization metrics we are aware of can be expressed as averages of some target functions. For example, the density of infected computers with a virus in the Internet, the ratio of sport websites in the WWW networks, and the average number of copies of a file in a Peer-to-Peer network, can be computed by the average of the target function $f$: $f(x)=1$, if node $x$ has desired characteristic (respectively, is infected with that virus, is a sport website, has a copy of that file), and $f(x)=0$, otherwise.
A measurement framework for computing the average of such functions can be achieved in two steps; (1) Sampling, and (2) Estimation. In many cases (e.g. the Internet), the global structure of a network is initially unknown and there is no sampling frame (i.e., a list of all nodes within a network, from which nodes can be directly sampled). Therefore, using classical sampling methods such as random sampling are either impossible or impractical, and link-tracing based sampling methods are the only feasible solutions to collect data from such networks. In these methods, one can see the neighbors of already sampled nodes and make a decision on which nodes to visit next. Snowball (which is similar to breadth-first search in computer science) \cite{Frank2011} and Random walk based methods \cite{Heckathorn1997} are among the most popular link-tracing based methods.
In the Estimation step, an estimator is used to approximate the network characteristics. An estimator is a function that accepts a summary of the sampled data as input and outputs an estimate of an unknown network characteristic. Specifically, the selection bias of a sampling method can be corrected by using the general idea of Hansen-Hurwitz estimator \cite{Hansen1943}, i.e. the measured value is weighted inversely proportional to the visiting probability of the corresponding node \cite{Volz2008}.
Therefore, we need to compute (or approximate) the visiting probabilities of each sampled node. In particular, in undirected networks, when data from snowballs with more than one wave are available, it might be comparatively easy to compute the visiting probabilities of each sampled node based on the first few waves, but it is more complicated to find the required visiting probabilities for snowballs of many waves \cite{Frank2011, kurant2011towards}. For random walk based sampling methods, in a sufficiently long ergodic walk on an undirected network, we would expect that the probability that a node is visited is proportional to its degree \cite{Lovasz1993}.
Many of the networks around us such as WWW, Twitter, and Peer-to-Peer networks contain directed links, or links which do not have the same strength in each direction. However, despite recent efforts on characterizing undirected networks based on the sampled data \cite{Salehi2012, Gjoka2011a, Gjoka2011b, Ribeiro2010}, little attention has been made on the statistical properties of sampled directed networks \cite{Lu2011, Ribeiro2012}. The directed links pose a considerable challenge to the process of network measurement (i.e., unknown bias in the sampled nodes). In summary, computing (or approximating) the visiting probabilities of sampled nodes for directed networks still remains as an open problem. For example, the stationary distribution for random walks on a directed network is no longer determined by the degree sequences. We can only show that a random walk process on a strongly connected directed network has a unique stationary distribution (but, there is no analytical solution for it yet).
However, most real-world directed networks are not strongly connected. Without this property, the limiting distribution of a random walk is sensitive to the starting distribution, and its analysis can be very complicated. A modification of random walk which overcomes this problem allows for a jump with small probability to a node at random from the whole network. That is called PageRank in the Internet search literature \cite{Brin1998}. It's clear that according to our assumptions about the absence of sampling frame, jumping to a random node from the whole network may not be possible. Therefore, PageRank can not be applied to our problem in its global definition.
In this paper, we propose a complete framework to measure nodal characteristics which can be defined by average target functions in an arbitrary directed network. In the sampling step, we introduced a novel link-tracing algorithm by utilizing the general idea of PageRank. We employ a modified version of PageRank, called personalized PageRank \cite{jeh2003}, using a specific set of nodes as a seed vector. Jumping is limited to this seed vector. Thus, personalized PageRank determines the importance of every node to the seed by using local information. In particular, the proposed method samples the underlying network by moving from a node to one of its neighbors through an outgoing link, based on the approximated probability of personalized PageRank.
Since these probabilities represent an approximation of the exact visiting probabilities, we utilize the idea of approximate ratio importance sampling to propose a new estimator.
The previous well-known work \cite{Volz2008, Salganik2004} on the estimators for correcting the bias, does not consider the effect of approximation of visiting probabilities which we address in this paper.
Importance sampling is one of the methods of Monte Carlo simulation \cite{Liu2001} in which instead of generating a sample from the target distribution (e.g., uniform distribution in our case), the estimator generates a sample from a different trial distribution (e.g., the distribution generated by the proposed sampling algorithm). We evaluate our framework over large synthetic and real datasets in terms of estimated average of a target function (here, we focus on the infection ratio and outdegree distribution) and its bias. Our results show that the proposed method performs well even in low sampling rates.
\section{Related Work}
\label{sec:Related Work}
Link-tracing based sampling methods in the context of networks are somewhat distinct from classical sampling methods \cite{Frank2011}. Random walk (RW) \cite{Lovasz1993} is one of the most important and widely used sampling methods in different kind of network contexts such as uniformly sampling Web pages from the Internet \cite{Henzinger2000}, content density in Peer-to-Peer networks \cite{Stutzbach2008}, degree distributions of the Facebook social graph \cite{Gjoka2011b}, and collecting data from information diffusion networks \cite{Eslami2012a, Eslami2012b}. A classic RW, samples a graph by moving from a node, $x$, to a neighboring node, $y$, through an outgoing link, chosen uniformly at random from the neighbors of $x$. By this process links and nodes are sampled.
In any given connected and non-bipartite undirected graph, the classic RW is biased towards node with higher degree.
In general, the probability of selecting the next node determines the probability that nodes are visited in sampling procedure. Metropolis-Hastings technique \cite{Metropolis1953} can be used to modify the probabilities of the node selection in random walk in order to have the uniform stationary distribution for visiting each node. This technique is a general Markov Chain Monte Carlo (MCMC) method \cite{Chib1995} for sampling from probability distributions based on constructing a Markov Chain that has the desired distribution as its stationary distribution. This approach, known as Metropolis-Hastings Random Walk (MHRW), has been applied to Peer-to-Peer networks \cite{Stutzbach2008}, Facebook \cite{Gjoka2011b}, and Twitter \cite{Salehi2011}.
Alternatively, one can use the unmodified classic RW method to sample from a graph and correct the degree bias by re-weighting the sampled values. The Respondent-driven sampling (RDS) \cite{Heckathorn1997,Salganik2004} is an example of this approach. Actually, this is a framework in the field of social sciences to sample and infer in hard-to-reach populations such as injection drug users. In these populations, a sampling frame for the target population is not available. Sampling from OSNs or WWW network graph is analogous to the sampling of hidden population in the social sciences. In the context of graph sampling, the RDS method has been used for Facebook \cite{Gjoka2011b}, Twitter \cite{Salehi2011}, and Peer-to-Peer networks \cite{Stutzbach2008}.
One of the main assumptions of the above mentioned studies is that the underlying network is undirected. Since most real-world networks (such as WWW, Twitter, and Peer-to-Peer networks) are directed, this assumption is not easily met in real life. However, little attention has been paid to directed networks by considering the effect of directed links. The authors in \cite{Hall2009} consider directed links in the Peer-to-Peer network while maintaining desirable statistical properties (i.e., uniformity) for the sampling procedure. The main idea of this RW based method is to avoid the calculation of each node's visiting probability, and instead to adjust the transition probabilities iteratively, converging to a state in which the sum of transition probabilities into each node equals $1$. The resulting transition matrix is said to be doubly stochastic, and induces uniform stationary distribution.
In \cite{Lu2011}, it is shown that when the sample size is relatively small, the RDS method may generate relatively large biases and errors if the studied networks are directed, indicating that estimates from previous RDS studies should be interpreted and generalized with caution. In \cite{Ribeiro2012}, a random walk sampling algorithm with jumps, called Directed Unbiased Random Walk (DURW), is proposed that achieves asymptotically unbiased estimates of the outdegree distribution of a directed graph. The authors construct an undirected graph using the nodes that are selected by the random walker on the directed graph. The undirected graph is built to allow the walker to traverse known outgoing links backwards, and guarantees that the probability of sampling a node can be approximated, even though incoming links are not observed. However, this method is based on the assumption that nodes can be sampled uniformly at random from the original graph (i.e., there is some means of obtaining a list of nodes in the network) which is not always feasible. For example, while it is feasible for Wikipedia and Twitter, it is not feasible for the WWW graph.
In conclusion, while the effect of directed links have been studied, there is no efficient measurement framework for directed networks in the literature. Moreover, no analytical solution for the visiting probabilities of the nodes is available for a general directed network.
\section{Preliminaries}
\label{sec:Preliminaries}
\subsection{Basic Notations and Definitions}
Let $G = (V, E)$ with $n=|V|$ and $m=|E|$ be the graph representing a complex network, where $V$ is the set of nodes, and $E=\{(i,j): i,j\in V, A_{ij}=1\}$ is the set of unweighted directed links between pairs of nodes. $A_{ij}$ indicates the presence ($=1$) or absence ($=0$) of relation of interest (such as a hyper-link in WWW, friendship in a social network, and file transferring in a Peer-to-Peer network) from $i$ to $j$.
The above $n \times n$ matrix $A$ is called an adjacency matrix.
The indegree of a node $x$, $d_{in}(x)$, is the number of distinct ingoing links $(y_1 , x), . . . , (y_k , x)$ into $x$, and its outdegree, $d_{out}(x)$, is the number of distinct outgoing links $(x, y_1 ), . . . , (x, y_k)$ out of $x$.
A link is called reciprocal if a connected pair of nodes have both an ingoing and an outgoing link between each other.
The proportion of reciprocal links in a network called reciprocity, $r$ (i.e. $r = 1$ when the network is undirected, and $r = 0$ when the network is directed in a way such that there are no reciprocal links).
Let $S$ be a sample of nodes where $S \subset V$, and $G(S)$ is the induced subgraph of $G$ based on the sample $S$.
Let $N(S)$ denote the neighborhood of $S$;
\begin{equation} \label{NeighborhoodEqu}
N(S) = \{w \in V \-- S: \exists v \in S \hspace{2mm} s.t. \hspace{2mm} (v,w) \in E\}
\end{equation}
Therefore, $N(S)$ is the set of nodes that we know exist due to other outgoing links but have not yet visited.
Let $L$ be a set of real-valued labels. For instance, a label can be the degree of a node. We assume that a label $l_x \in L$ is assigned to each node $x \in V$ by a target function $f:V \rightarrow L$, i.e. $f \subseteq \{(x,l_x): x\in V, l_x \in L\}$. We can express many network characterization metrics as averages of some target functions over $V$. Given a target function $f$, the average of $f$ is: $avg(f)={\sum_{x \in V} f(x)/ |V|}$.
Our main goal in this paper is to propose a measurement framework to compute $avg(f)$ (based on the sampled nodes $S$) for a general target function $f$ (that defines a nodal characteristic) in an arbitrary directed network $G$. Our only assumption on $G$ is that it is weakly connected, i.e. every node can be reached from every other node by traversing the links without considering the direction of the links.
\subsection{Personalized PageRank Vectors}
PageRank was first introduced by Brin and Page \cite{Brin1998} for search algorithms in the WWW network. However, it can be defined for any graph as a stationary distribution of a certain random walk over that graph. At each step, with probability $(1-\alpha)$, the random walk follows a randomly outgoing link of a node, and with probability $\alpha$ the random walk makes a jump to a new node chosen uniformly among all nodes in the network. The jumping constant $\alpha$ ($0 < \alpha \leq1$), controls the diffusion rate. With smaller $\alpha$, the random walks spread further from the initial nodes before performing a random jump. Traditionally, the value of $\alpha$ is chosen to be $0.15$.
Let $A$ be the adjacency matrix of a graph $G$. The PageRank vector $\mathbf{p(\mathbf{s})}$ of this graph is a visiting probability distribution on the nodes of $G$ that can be defined as the solution of the following equation:
\begin{equation}
\label{PageRankEqu}
{\mathbf{p(\mathbf{s})}} = \alpha{\mathbf{s}}+(1-\alpha){\mathbf{p} D_{A}^{-1}A}
\end{equation}
where $\mathbf{s}$ is a seed vector that includes the initial distribution for starting nodes.
We use the notation $D_{A}^{-1}$ to indicate the diagonal matrix with the inverse outdegree for each node. Globally, uniform $\mathbf{s}=\mathbf{1}/n$ is considered for computing PageRank. However, an arbitrary distribution ${\mathbf{s}}$ can lead to creation of personalized PageRank \cite{Haveliwala2003}. Specifically, the random walk in the personalized PageRank only jump to a few nodes of personal interest. It can be shown that for any ${\mathbf{s}}$ and $\alpha$, there is an unique vector $\mathbf{p(\mathbf{s})}$ satisfying Equ.(\ref{PageRankEqu}) \cite{jeh2003}.
There are various algorithms for computing global PageRank and personalized PageRank \cite {jeh2003, berkhin2006, gleich2006}. As a standard PageRank algorithm, we can efficiently compute the solution of Equ.(\ref{PageRankEqu}) by applying the power method \cite{Golub1996}.
Given some initial distribution $\mathbf{s}$, the power method is defined as following iteration:
\begin{center}
${{\mathbf{p}^{(k+1)}} = (1-\alpha) {\mathbf{p}^{(k)}} D_{A}^{-1} A}$
\\
${{\mathbf{p}^{(k+1)}} = {\mathbf{p}^{(k)}} + (1- {\| {\mathbf{p}}^{(k+1)} \|_{1}})\mathbf{s}}$
\end{center}
where $ {\| .\|_{1}}$ denotes the $1$-norm. As an algorithm, the power method continues this iteration until ${\|{\mathbf{p}^{(k+1)}} - {\mathbf{p}^{(k+1)}}\|_{1}} < \delta$ for a user-provided stopping tolerance $\delta$.
Based on the idea of the power method, Gleich and Polito \cite{gleich2006} present an efficient algorithm, called boundary-restricted personalized PageRank algorithm (BRPPR), to compute an approximation of the personalized PageRank vector $\mathbf{p(\mathbf{s}})$ by examining only a small fraction of the input web graph near the starting vector $\mathbf{s}$. Specifically, this algorithm iteratively divide the web pages into an active and inactive set. At each iteration, the set of active web pages is expanded to include more web pages that are likely to have a high personalized PageRank value.
Specifically, it expands pages until the total rank on the frontier set (the set of pages that we know exist due to other outgoing links but have not yet visited) of pages is less than an expansion tolerance.
Therefore, only the set of active pages and their outgoing link information are actually involved in the computation. Moreover, a stopping tolerance threshold is considered in order to determine the termination of the approximation algorithm. The analysis of BRPPR leads to the following theorem (theorem 3.3 and remark 3.4 of \cite{gleich2006}):
\begin{mytheorem}
\label{Theorem:ApproximatePR}
The BRPPR algorithm with expansion tolerance $\kappa$ and stopping tolerance $\delta$ yields an approximate PageRank vector $\hat{\mathbf{p}}$, where $ {\| {\mathbf{p}} - \hat{\mathbf{p}}\|_{1}} \leq {{2(1-\alpha)}\over{\alpha}}. \kappa +{{2-\alpha}\over{\alpha^2}}.\delta $\\
\end{mytheorem}
where ${\| {\mathbf{p}} - \hat{\mathbf{p}}\|_{1}}$ gives the Manhattan distance between two distribution vectors of $\mathbf{p}$ and $\hat{\mathbf{p}}$. Although, the authors in \cite{gleich2006} address the problem of calculating the PageRank score of a Internet webpage in the field of search engines, we utilize their general idea here in the context of network sampling.
\subsection{Importance Sampling}
\label{sec:Approximate Importance Sampling}
As mentioned, we are interested in measurement of quantities that can be written as averages of some target functions $f$ over a finite set $V$ (i.e., the set of all nodes).
Let $\pi:V\rightarrow [0, \infty)$ be the target distribution on a set $V$. In our problem, the desired $\pi$ is the uniform distribution on $V$, then all nodes $x$ are equally likely to be selected, i.e. $\pi(x)=1/|V|$. Let $\hat{\pi}: V\rightarrow [0, \infty)$ be a target measure that is unnormalized form of $\pi$, i.e. $\hat{\pi}(x)={\pi(x)}.{Z_{\hat{\pi}}}$ where $Z_{\hat{\pi}}>0$ is the normalization constant of ${\hat{\pi}}$. We say that measure ${\hat{\pi}}$ induces a corresponding probability distribution on $V$.
Therefore, ${\hat{\pi}}$ is a relative weight, which represents the probability of $x$ to be chosen in the distribution $\pi$. In our problem, we have ${\hat{\pi}}=1$ (for all $x \in V$) and $Z_{\hat{\pi}} = |V|$.
The average of a target function relative to $\hat{\pi}$ is essentially a sum where the target measure is the probability distribution $\pi$. Hence, we can write
\begin{equation}
avg(f)=avg_{\hat{\pi}}(f)=sum_{{\pi}}(f)={\sum_{x\in V} f(x).{{\pi}(x)}}
\end{equation}
For example, in our previous example of infection in a population, the infection ratio is an average of the target function relative to an uniform target measure. Alternatively, it is a sum of the same function relative to the uniform distribution on $V$ (i.e., $\pi(x) = 1/|V|$ for all $x$).
The naive estimator for $avg_{\hat{\pi}}(f)$ works as follows \cite{Liu2001}: (1) Generate a random sample $X$ from the distribution $\pi$ induced by $\hat{\pi}$, (2) Compute the function of $f$ for sampled $X$ (i.e., $f(X)$).
It is easy to check that this estimator is unbiased. However, since we assume that there is no sampling frame in our problem, uniform sampling from the distribution $\pi$ may be hard or costly. Consequently, this simple estimator is inefficient. To solve this problem, we can use \textit{Importance Sampling (IS)}, as one of the methods of monte carlo simulation \cite{Liu2001}. In the following, we state some theorems about the general idea of importance sampling \cite{Gurevich2011}. Although, the authors in \cite{Gurevich2011} address the problem of externally measuring aggregate functions over documents indexed by search engines, we use them here in the context of network sampling.
\subsubsection{Importance Sampling (IS)}
The basic idea of importance sampling is: instead of generating a sample $Y$ from the target distribution $\pi$, the estimator generates a sample $X$ from a different trial distribution $p$ on $V$ to directly estimate statistical sums relative to the target measure $\hat{\pi}$. Specifically, given a target function $f$, importance sampling estimator estimates $sum_{\hat{\pi}}(f)={\sum_{x\in V} f(x).{\hat{\pi}(x)}}$.
The trial distribution $p$ can be any distribution, as long as $supp(\pi) \subseteq supp(p)$ (here, $supp(p)=\{x \in V | p(x)>0\}$ and $supp(\pi)$ is defined similarly). In particular, we can choose $p$ to be a distribution that is easy to sample from. By considering $\hat{p}$ as unnormalized form of $p$, the importance sampling estimator is then defined as follows:
\begin{equation}
E_{IS}(X)=f(X).{{\hat{\pi}(X)}\over{\hat{p}(X)}}=f(X).w(X), \hspace{5mm} X\in V
\end{equation}
where $w(X)$ is called the importance weight. Function $f$ is computed by samples of distribution $p$.
It can be shown that $E_{IS}(X)$ is an unbiased estimator for $sum_{\hat{\pi}}(f)$.
Implementation of an importance sampling estimator requires: (1) Ability to sample efficiently from the trial distribution $p$; and (2) Ability to compute the importance weight $w(x)$ (or its estimator) and the function value $f(x)$, for any given node $x \in V$. There is no need to know the normalization constant $Z_{\hat{\pi}}$ or to be able to sample from $\pi$. This basic importance sampling estimator is only suitable for estimating sums. We next extend it for estimating averages.
\textbf{Importance Sampling for average:}
Recall that the estimator for averages estimates $avg_{\hat{\pi}}(f) = sum_{\pi}(f)$. If we are able to compute $\pi(x)$ for each $x \in V$, we can use the importance sampling estimator described above to estimate average of a function relative to $\hat{\pi}$. However, in many cases we can only compute $\hat{\pi}(x)$ and not $\pi(x)$. For example, in our case, $\pi(x) = 1/|V|$, where $|V|$ is the number of nodes, that is typically unknown.
Ratio importance sampling is a technique for estimating averages relative to any target measure $\hat{\pi}$, even when $\pi(x)$ is unknown.
The ratio importance sampling estimator computes two sum estimators: $M_1$ of $sum_{\hat{\pi}}(f)$, and $M_2$ of $sum_{\hat{\pi}}(w)$, and then outputs $M_1/M_2$.
However, there is one problem:
the expectation of a ratio is not the ratio of the expectations, i.e., $\mathds{E}(M_1/M_2) \neq \mathds{E}(M_1)/\mathds{E}(M_2)$.
To solve this problem, ratio importance sampling utilize the following idea: if we replace $\mathds{E}(M_1)$ and $\mathds{E}(M_2)$ by averages of multiple independent instances of the estimator of $\mathds{E}(M_1)$ and $\mathds{E}(M_2)$, the difference between $\mathds{E}(M_1/M_2)$ and $\mathds{E}(M_1)/\mathds{E}(M_2)$ diminishes to $0$. Therefore, the ratio importance sampling estimator for $avg_{\hat{\pi}}(f)$ is then defined as follows:
\begin{equation}
E_{RIS}(X_1,X_2,...,X_n) = {{{1 \over n} {\sum\limits_{i=1}^n E_{IS}(X_i)}} \over {{1 \over n} {\sum\limits_{i=1}^n w(X_i)}}}
\end{equation}
where $X_1,X_2,...,X_n$ are $n$ independent samples from the trial distribution $p$.
\subsubsection{Approximate Importance Sampling (AIS)}
Unfortunately, in many cases we are unable to accurately compute the importance weight function $w$. This problem arises, e.g. in our case, due to complexity of analysing a RW based sampling method on a directed network. In the following, we address the effect of the approximate importance weights on importance sampling.
Approximate importance sampling method employ an \lq\lq{approximate importance weight function}\rq\rq $u(x)$ rather than the exact one $w(x)$. This estimator can be defined as:
\begin{equation}
E_{AIS}(X)=f(X).u(X)
\end{equation}
where $X$ is distributed according to the trial distribution $\hat{p}$. The estimation generated by approximate importance sampling is close to the true value as long as the importance weight function $w(x)$ and the approximate importance weight function $u(x)$ are similar. To analyse the bias of this estimator, we consider the following theorem.
\begin{mytheorem}
\label{Theorem:biasOfEST_AIS}
(\cite{Gurevich2011}) Suppose $supp(u) \subseteq supp(w)$\\
${\mathds{E}(E_{AIS}(X))}=sum_{\hat{\pi}}(f(Y)).\mathds{E}({u(Y) \over {w(Y)}})+Z_{\hat{\pi}}.cov(f(Y),{u(Y) \over w(Y)})$\\
\end{mytheorem}
where $X \sim {\hat{p}}$, $Y \sim \pi$, and $Z_{\hat{\pi}}$ is the normalization constant. Therefore, there are two types of bias in this estimator: (1) multiplicative bias, depending on the expectation of $u/w$ relative to $\pi$. It may have a significant effect on the estimator's bias, and thus must be removed. (2) additive bias, depending on the correlation between $f$ and $u/w$ and on the normalization constant $Z_{\hat{\pi}}$. It is typically less significant, as in many practical situations $f$ and $u/w$ are uncorrelated (e.g., when $f$ is a constant function, as is the case with infection ratio example).
To remove the multiplicative bias, we assume that it is possible to estimate the multiplicative bias, and divide $E_{AIS}(X)$ by this estimate. For an instance $x \in V$, the ratio $u(x)/w(x)$ is called the weight skew at $x$. The multiplicative bias factor is the expected weight skew relative to the target distribution $\pi$.
To remove this bias, we need to estimate the expected weight skew. Let $E_{WSE}(X)$ be an unbiased weight skew estimator for $\mathds{E}({{u(Y)} \over {w(Y)}})$.
It follows from Theorem \ref{Theorem:biasOfEST_AIS} that:
\begin{equation}
{{\mathds{E}(E_{AIS}(X))} \over {\mathds{E}(E_{WSE}(X))}}=sum_{\hat{\pi}}(f(Y))+{Z_{\hat{\pi}}.{cov(f(Y),{u(Y) \over w(Y)}} \over {\mathds{E}({u(Y) \over {w(Y)}})}}
\end{equation}
Thus, the ratio of the expectations of the two estimators, $EST_{AIS}(X)$ and $EST_{WSE}(X)$, gives us the desired result $sum_{\hat{\pi}}(f)$, modulo an additive bias factor that depends on the correlation between $f$ and $u/w$. The additive bias can be considered $0$ since $f$ is often a constant function. Ignoring this additive bias, it would seem that the ratio ${EST_{AIS}(X)} / {EST_{WSE}(X)}$ is a good estimator for $sum_{\hat{\pi}}(f)$.
Although the expectation of a ratio is not the ratio of the expectations, one applies the ratio estimation technique similarly to what we did in the ratio importance sampling.
Therefore, the approximate ratio importance sampling estimator for $sum_{\hat{\pi}}(f)$ can be defined as follows:
\begin{equation}
E_{ARIS}(X_1,X_2,...,X_n) = {{{1 \over n} {\sum\limits_{i=1}^n E_{AIS}(X_i)}} \over {{1 \over n} {\sum\limits_{i=1}^n E_{WSE}(X_i)}}}
\end{equation}
where $X_1,...,X_n$ are $n$ independent samples from the trial distribution $\hat{p}$.
\textbf{Approximate importance sampling for averages:}
Here, we employ $E_{ARIS}$ to design an importance sampling estimator for averages. Recall that the estimator for averages estimates $avg_{\hat{\pi}}(f) = sum_{\pi}(f)$. Thus, we need to compute the importance weight function $w_{avg}(x)={\pi}(x)/{\hat{p}}(x)$.
However, $\pi(x)$ cannot be calculated exactly. For solving this problem, we use an approximate weight function $u(x) \approx w_{avg}(x)$ instead of $w_{avg}(x)$, and then fixing the bias using approximate ratio importance sampling estimator (i.e., $E_{ARIS}$) described above. To this end, we need to come up with the weight skew estimator whose expectation equals $\mathds{E}(u(Y)/w_{avg}(Y))$, where $Y$ is distributed according to $\pi$. It can be shown that with a constant function $f$, the approximate weight $u(X)$ itself is an unbiased estimator for $\mathds{E}(u(Y)/w_{avg(Y)})$ \cite{Gurevich2011} . We thus set $E_{WSE}(X) = u(X)$. Therefore, approximate importance sampling for averages, $E_{AVG}$, is then defined as follows:
\begin{equation}
\label{equ:AIS_avg}
E_{AVG} = {{{1 \over n} {\sum\limits_{i=1}^n f(X_i).u(X_i)}} \over {{1 \over n} {\sum\limits_{i=1}^n u(X_i)}}}
\end{equation}
where $X_1,...,X_n$ are $n$ independent samples from the trial distribution $\hat{p}$. The following theorem shows the bias and the variance of this estimator.
\begin{mytheorem}
\label{Theorem:biasOfE_AVG}
(\cite{Gurevich2011}) The bias and variance of $E_{AVG} $ (equ. \ref{equ:AIS_avg}) are defined as
\\
$Bias(E_{AVG})=E_{AVG}(X_1,X_2,...,X_n) - avg_{\hat{\pi}}(f)={Z_{\hat{\pi}}.{cov(f(Y),{u(Y) \over w_{avg}(Y)})} \over {\mathds{E}({u(Y) \over {w_{avg}(Y)}})}} + {O({1 \over n})}$
\\
$var(E_{AVG}(X_1,X_2,...,X_n))={O({1 \over n})}$
\end{mytheorem}
\section{Proposed Measurement Framework}
\label{sec:ProposedMeasurmentProcess}
The pseudo code of proposed framework, called Directed-Network-Measurement (DNM), is shown in Algorithm \ref{DNM-algorithm}.
Our main goal is to approximate the probability of each visited node in sampling procedure, and correct the bias of sampling by using the idea of approximate importance sampling.
The proposed framework is designed through two steps; sampling and estimation. The sampling step initially starts with an empty sample set $S$, and adds a random node $y$ to it as a seed. Subsequent nodes of the sample are chosen based on the following procedure. In each iteration, the set of neighborhood of already sampled nodes, i.e. $N(S)$, and its corresponding active link matrix, $L$, is generated. The matrix $L$ is the adjacency matrix for all outgoing links from nodes in $S$, but may reference nodes that are not in $S$. We then compute one PageRank iteration on $L$ from a vector $\hat{\mathbf{p}}$ to a vector $\tilde{\mathbf{p}}$. Next, we add nodes (that are likely to have a high personalized PageRank value) to $S$ until the total probability on the remainder of the neighborhood is less than an expansion ratio of $\kappa$.
As we mentioned earlier, the essential property of a sampling method that makes it appropriate for network inference is that its visiting probabilities are known or estimable for the sampled nodes. In the estimation step, we use computed PageRank value of each sampled node (i.e. $\hat{\mathbf{p}}$), as an approximation of the exact visiting probability. Then, we utilize the idea of approximate importance sampling to design a new estimator and study its performance for a general target function. Specifically, we consider the obtained $\hat{\mathbf{p}}(x)$ as an approximate weight function $u(x)$.
According to the general ratio importance sampling estimator (refer to Equ.\ref{equ:AIS_avg}), we propose the following estimator for computing the average of a general target function $f$ in a directed network.
\begin{equation}
\label{est:dir}
E_{DIR} = {{{1 \over n} {\sum\limits_{i=1}^n f(X_i).\hat{\mathbf{p}}(X_i)}} \over {{1 \over n} {\sum\limits_{i=1}^n \hat{\mathbf{p}}(X_i)}}}
\end{equation}
where $X_1, X_2, ..., X_n$ are $n$ samples extracted by sampling procedure.
\begin{pseudocode}[ruled]{DNM}{y, \alpha, \kappa, \delta} \label{DNM-algorithm}
\mbox{$S = \{y\}$; $L$ = matrix of $y$ to its outgoing links}\\
\mbox{SUM = $0$}\\
\mbox{WSE = $0$}\\
\mbox{$\hat{\mathbf{p}}=0; \hat{\mathbf{p}}(y) = 1$}\\
\mbox{\% Sampling step} \\
\REPEAT \DO
\BEGIN
\mbox{$\tilde{\mathbf{p}} = (1-\alpha) \hat{\mathbf{p}} D_{L}^{-1} L$}\\
\mbox{$\tilde{\mathbf{p}} = \tilde{\mathbf{p}} + (1- {\| \tilde{\mathbf{p}} \|_{1}}$)}\\
\WHILE {\tilde{\mathbf{p}}(N(S)) > \kappa} \DO
\BEGIN
\mbox{Find $x \in N(S)$ with max value in $\tilde{\mathbf{p}}$}\\
\mbox{Add $x$ to $S$}\\
\mbox{Remove $x$ from $N(S)$}\\
\mbox{Update $\tilde{\mathbf{p}}(N(S))$}\\
\END\\
\mbox{Update $D_L$ and $L$}\\
\mbox{$\omega = \|\tilde{\mathbf{p}} - \hat{\mathbf{p}}\|_{1}$}\\
\mbox{$\hat{\mathbf{p}} = \tilde{\mathbf{p}}$}\\
\END\\
\UNTIL{\omega < \delta}\\
\mbox{\% Estimation step} \\
\FOR {x \in S} \DO
\BEGIN
\mbox{SUM = SUM + $f(x).\hat{\mathbf{p}}(x)$}\\
\mbox{WSE = WSE + $\hat{\mathbf{p}}(x)$}\\
\END\\
\RETURN{\mbox{SUM/WSE}}
\end{pseudocode}
The following theorem shows the bias and variance of the proposed estimator.
\begin{mytheorem}
\label{Theorem:estDNM}
The bias and variance of $E_{DIR} $ (equ. \ref{est:dir}) are defined as\\
$Bias(E_{DIR})=E_{DIR}(X_1,X_2,...,X_n) - avg_{\hat{\pi}}(f)={{|V|}.{cov(f(Y),{{\hat{\mathbf{p}}} / {\mathbf{p}}})} \over {\mathds{E}({{\hat{\mathbf{p}}} / {{\mathbf{p}}}})}} + {O({1 \over n})}$
\\
$var(E_{DIR}(X_1,X_2,...,X_n))={O({1 \over n})}$\\
\end{mytheorem}
where $Y \sim \pi=1/|V|$, and $Z_{\hat{\pi}}$ is the normalization constant.
\textbf{proof:}
By substituting $u=1/|V|$, $w_{avg}={\mathbf{p}}$, and $Z_{\hat{\pi}}=|V|$ into Theorem \ref{Theorem:biasOfE_AVG}, the theorem is proved.
We conclude from this theorem that if we use sufficiently large number of samples, then we are likely to obtain an estimate of $avg_{\hat{\pi}}(f)$, which has only additive bias that depends on the correlation between $f$ and ${{\hat{\mathbf{p}}} / {{\mathbf{p}}}}$.
\section{Experimental Evaluation}
\label{sec:expEval}
\subsection{Datasets}
\label{subsec:Datasets}
\textbf{Synthetic Networks:} We utilize three kinds of random models for generating directed networks:
Directed Erdos-Renyi (DER) networks \cite{Erdos59}: The model starts with $|V|$ nodes, which link to each other with probability $p_0 \in (0, 1)$. The reciprocity of these network equals to $1$. In order to decrease the reciprocity, reciprocal links in the base network are randomly chosen and converted to irreciprocal links. We generate networks by using this model with $|V|=2000$, $p_0=0.1$, and $r=0.6$.
Directed Watts-Strogatz (DWS) networks \cite{Watts98}: The model starts from a completely regular network with identical degree ($k$-nearest neighbors' connection) and clockwise links. Each link will be rewired with two randomly selected nodes with probability $p_0 \in (0, 1)$. In simulation, we generate networks with $|V|=2000$ nodes by setting $k=20$ and $p_0=0.1$.
Directed Scale Free (DSF) \cite{Pennock2002}:
Network starts with $m_0$ nodes, which link to each other with probability $p_0$ (as in an Erdos-Renyi random graph). At each time step, one node and $m$ links are added to the network. The endpoints of the links are randomly selected among all nodes according to the following probability: $p_1(x) = \beta_1 {{d_{in}(x)}\over{|E|}}+ \beta_2 {{d_{out}(x)}\over{|E|}}+\beta_3 {{1}\over{|V|}}$, where $\beta_1 + \beta_2 + \beta_3 = 1$. We generate networks with $|V|=1000$ and average degree of nodes $=50$. We leave the number of links, $|E|$, unconstrained. For attaching probability, we set $\beta_1=0.7, \beta_2=0.2, \beta_3 = 0.1$. Moreover, we set $m_0=2$ for generating initial Erdos-Renyi graph and consider $p_0=1$ to have them fully connected.
\textbf{Real-world Networks:} Moreover, we considered some real-world directed network:
An Online Social Network (OSN): This is a Facebook-like Social Network \cite{Opsahl2009} that originate from an online community for students at University of California, Irvine. The dataset includes the 1,899 users that sent or received at least one message. There are 20,296 directed links among these users.
A Peer-to-Peer Network (P2P): We use a snapshot of the Gnutella Peer-to-Peer file sharing network collected in August 2002 \cite{Leskovec2007a}. Nodes and links represent hosts and connections between them in this network, respectively. The dataset includes the 10,876 hosts and 39,994 directed links.
The Internet Autonomous System (AS): The network of routers comprising the Internet can be organized into sub-networks called Autonomous Systems (AS). Each AS exchanges traffic flows with some neighbors. We use a CAIDA AS graph collected in February 2004 \cite{Leskovec2005}. The dataset contains 16,493 nodes and 66,744 links.
\subsection{Target Function}
\label{subsec:targetFuncion}
The fast growth of Internet and other communication networks makes them a suitable target for malicious activities. An infection spreads through the links of such networks constructed by computers and their communication channels. The process by which malicious objects such as worms, trojan horses, and computer viruses travel through computer networks is analogous to the process of spreading epidemics through a population. In this paper, we study the infection ratio in spreading of a disease in a population as a network characteristic. It is an example of family of binary properties of a network where each node is tagged by label $0$ or $1$. In particular, infection ratio can be considered as the average of a target function $f$: $f(x)=1$, if individual $x$ is infected, and $f(x)=0$, otherwise.
To spread disease on the test datasets, we use the susceptible-infectious-recovery (SIR) epidemic spreading model \cite{Moreno2002}. This is an epidemiological model widely used to simulate the spreading of epidemics, i.e. number of people infected with a contagious disease, in a population as a function of time. At every time step, the model assumes a transition rate $\theta_1$ for a susceptible person to become infected, if an infected neighbor is present, and a rate $\theta_2$ for an infected person to become recovered or die. The recovered person will never be infected again. In the simulations, we use $\theta_1=0.2$ and $\theta_2=0.05$. We run the simulation until desired ratio of infected nodes obtained.
\subsection{Experimental Setups}
\label{subsec:Experimental Setups}
We evaluate the performance of the DNM framework in various situations and in terms of different aspects. We consider two metrics; $1)$ estimated infection ratio, and $2)$ bias (which is defined by the absolute difference between estimated and true infection ratio).
For each sampling rates in the range of $0.01\%$ to $40\%$, we repeated the simulations 100 times by selecting random nodes as seed, and finally, averaged the obtained values. We set jumping factor $\alpha=0.15$ and stopping tolerance $\delta=10^{-7}$ as the inputs of the proposed framework.
Moreover, we set expansion tolerance $\kappa$ based on the desired sampling rates. Since, the parameter $\kappa$ controls the accuracy of the approximation of PageRank values (i.e., the visiting probabilities of the sampled nodes), different values of $\kappa$ provides different sampling rates. Specifically, the less $\kappa$ leads to more samples, and consequently the more accuracy is obtained. We study the obtained $\kappa$ per various sampling rates for all test networks. Our result is demonstrated in Figure $1$.
As we can see, the desired sampling rate depends on the structure of underlying network. According to Theorem \ref{Theorem:ApproximatePR}, the maximum error of the approximating visiting probabilities is in the range of $(0,6.8)$. This means that the proposed sampling method computes the visiting probabilities of the sampled nodes with reasonable accuracy even in low and medium sampling rates.
\begin{figure}[tp]
\begin{center}
\includegraphics[scale=0.29]{kappa_samples}
\end{center}
\caption {The relation between expansion tolerance ($\kappa$) of the proposed sampling method and sampling rate for all test networks.}
\label{fig:spk}
\end{figure}
\subsection{Evaluation Results}
\label{subsec:Evaluation Results}
\subsubsection{Main Results}
\label{results}
The results of our study on the performance of proposed framework is shown in Figure \ref{AE-main}. As we can see, in general, the error in terms of estimated infection ratio and bias decreases significantly by increasing the sampled nodes. In particular, we reach desirable result in medium sample rates (estimated infection ratio and bias converges to $20\%$ and $0$, respectively).
However, we found different behaviour in DWS networks; the estimated infection ratio is close to true value in all sampling rates (i.e. it's independent of number of samples). Moreover, the bias decreases with increasing the sampling rate.
\begin{figure}[tp]
\begin{center}
\subfigure[]{\label{AE-label}\includegraphics[scale=0.29]{AE_main}}
\\
\subfigure[]{\label{Bias-label}\includegraphics[scale=0.29]{Bias_main}}
\end{center}
\caption{The performance of the proposed framework in terms of (a) Estimated Infection Ratio (the true value is $20\%$) and (b) Bias.}
\label{AE-main}
\end{figure}
\subsubsection{Convergence analysis}
Deriving valid population estimates from sampled nodes is based on the assumption that the samples are derived from the equilibrium distribution, which is true asymptotically. In this section, we study the convergence of the DNM estimates to equilibrium distribution during the data collection step.
In general, the starting nodes (the seeds) are not selected from a sampling frame, but instead are ad-hoc samples. One way to reduce the dependence of final estimates to seed nodes is to use a burn-in period by discarding large numbers of initial sampled nodes before analyzing the collected data.
Given the high cost in terms of time and effort of collecting data, this may not be a desirable approach. Moreover, the authors in \cite{Gile2010} demonstrate that in a without-replacement sampling setting, this approach can even introduce more bias. In fact, the only real way to apply a burn-in to ensure accuracy would be to repeat the burn-in after every sampled node.
An alternative approach would be to estimate the relative visiting probabilities of all sampled nodes, conditioned on the composition of the seeds, and compute the estimates based on those probabilities. We follow this approach in the proposed framework. In particular, since jumping in personalized PageRank is limited to the seed vector, we approximate the importance of every node to the seed, i.e. the visiting probability, which is used in the Estimation step to correct the bias to the seeds.
To monitor the convergence of the proposed framework, we use a standard diagnostic test developed within the MCMC literature, namely Geweke \cite{Geweke1992}. This test was applied for the first time in the context of network sampling in \cite{Gjoka2011b}. The Geweke diagnostic detects the convergence of a single Markov chain by comparing the location of the sampled parameter on two different time intervals of the chain. The test is a standard Z-score with the standard errors adjusted for autocorrelation. We used this diagnostic in different runs of the proposed framework (generated by selecting some random nodes as seed) for the OSN dataset, and compares the difference between the first $10\%$ and the last $50\%$ of the samples. Figure \ref{fig:geweke1} presents the results of the Geweke diagnostic for the infection ratio as a network characteristic. We observe that after sampling approximately $200$ nodes, the $Z$-scores are strictly between $[-1, 1]$. This indicates that the proposed framework has achieved a good mixing with our initial selection of the random seeds.
\begin{figure}[htp]
\begin{center}
\includegraphics[scale=0.29]{geweke_OSN}
\end{center}
\caption {The Geweke convergence diagnostic of DNM for infection ratio in the OSN dataset. The lines show the Geweke $Z$-score for some runs with different random nodes as seed.
\label{fig:geweke1}
\end{figure}
\subsubsection{The effect of network reciprocity}
\label{dir}
Here, we study the performance of the proposed framework in networks with different levels of reciprocity $r$ which is an important parameter in characterization of directed networks. To this end, we generate some underlying directed networks with $r \in \{0.6, 0.7, 0.8, 0.9\}$ by DER model. As we can see the results in Figure \ref{Dir-whole}, in lower sampling rates, the amount of error in estimation of infection ratio decreases with decreasing reciprocity (we observe the same pattern in terms of bias). As a general conclusion, if we have to sample a few nodes from a network, lower proportion of reciprocal links in that network leads to more accurate outputs.
\begin{figure}[htp]
\begin{center}
\includegraphics[scale=0.29]{AE_recip}
\end{center}
\caption{The effect of reciprocity ($r$) on the performance of DNM in underlying networks generated by DER model. True value of infection ratio (= $0.2$) shown in the grey dotted line.}
\label{Dir-whole}
\end{figure}
\subsubsection{Sensitivity to true infection ratio}
\label{inf}
Figure \ref{Inf-whole} shows the sensitivity of DNM to infection ratio in DER, DWS, DSF, and OSN datasets (due to page limitation, we did not demonstrate the results for P2P and AS networks). We run the simulation of disease spreading until the true ratio of infected nodes reaches to desired values (specifically, $20\%$, $30\%$, $40\%$, and $50\%$). These ratios are used as the ground truth to performance evaluation.
According to the results, one can categorize the networks into three groups. The first group includes the networks generated by the DER model. In this group, the estimated infection ratio is more farther from the true value by increasing the true infection ratio. This error converge to zero for higher sampling rate in various infection ratios (In terms of bias, we observe the same pattern; higher infection ratio leads to higher bias in lower sampling rates). In second group, that includes the DWS networks, the infection ratio dose not have major impact on the estimated values and their biases.
\begin{figure*}[tp]
\begin{center}
\subfigure[DER]{\label{DER-AE}\includegraphics[scale=0.27]{AE_inf_DER}}
\subfigure[DWS]{\label{DWS-AE}\includegraphics[scale=0.28]{AE_inf_DWS}}
\\
\subfigure[DSF]{\label{DSF-AE}\includegraphics[scale=0.27]{AE_inf_DSF}}
\subfigure[OSN]{\label{OSN-AE}\includegraphics[scale=0.28]{AE_inf_OSN}}
\end{center}
\caption{The performance of the DNM framework in the networks with different infection ratios; (a) DER, (b) DWS, (c) DSF, and (d) OSN.}
\label{Inf-whole}
\end{figure*}
We found an inverse pattern in third group which is composed of the DSF and OSN (we observe the same results in the P2P and AS networks). In particular, lower true infection ratio leads to higher error in estimation of infection ratio (similarly, for obtained bias) for lower sampling rates. However, this error reduces by increasing the number of samples. The deduction rate is higher for the populations with lower infection ratio. As a general conclusion, in a scale free network with higher infection ratio, lower sampling rates are sufficient and result in a reasonable bias. More specifically, increasing the sampling rate does not have significant impact on the bias.
\subsubsection{Estimating Outdegree Distributions}
In what follows we compare the performance of DNM against the DURW sampling algorithm proposed by Ribeiro et al. \cite{Ribeiro2012} (presented in Section \ref{sec:Related Work}).
This is the most closely related method to ours that achieves asymptotically unbiased estimates of the outdegree distribution of a directed graph. However, there is a fundamental difference between the proposed framework and DURW; DNM only uses already visited nodes as local information without any prior knowledge about the latent structure of the network. In addition, the DURW algorithm is based on the assumption that nodes can be sampled uniformly at random from the original graph, which is not always feasible.
There are two controlling parameters in DURW. The first one is denoted by $c$, which is the cost of a random jump (i.e. the average number of sampling steps required to perform the jump).
The second one, $w$, is the random jump weight that controls the probability of performing a random jump.
Figure \ref{fig:outDegree} shows the comparison between estimates of the bias (in log scale) of DNM and DURW for all outdegrees in the OSN dataset (we observed the same results for other test networks). The bias was obtained over 100 runs. In all simulations we sampled $10\%$ of the network. The DURW random jump weight and cost was $w = 10$ and $c \in \{1,10,50\}$, respectively.
We found that both methods obtain accurate estimates. However, the fraction of nodes with large outdegrees can be estimated more accurately than the ones with small outdegrees. This is because both of them tend to sample nodes with larger outgoing links more frequently; sometimes causing lower estimation errors for the large outdegree nodes.
Moreover, as demonstrated in Figure \ref{fig:outDegree}, DURW outperforms DNM in terms of bias when the cost of a random jump is not significant. The cost of a jump effectively reduces the number of total nodes that can be sampled, which increases the bias. Therefore, we observe that the bias of the outdegree distribution estimates in DURW increases with $c$, and the DNM method is more accurate than DURW for larger costs.
\begin{figure}[h]
\begin{center}
\includegraphics[scale=0.29]{outDegree_OSN}
\end{center}
\caption{Bias of DNM compared with DURW ($w =10$ and $c \in \{1, 10, 50\}$) for all outdegrees in OSN dataset for sampling rate$=10\%$.}
\label{fig:outDegree}
\end{figure}
\section{Conclusion}
\label{sec:Conclusion}
In this paper, we introduced a novel two-step framework to measure nodal characteristics of large scale directed networks which can be defined by average target functions.
We proposed a novel link-tracing network sampling algorithm by utilizing the idea of personalized PageRank.
In particular, this method samples the underlying network by moving from a node to one of its neighbors through an outgoing link based on the approximated probability of Personalized PageRank.
Since these probabilities can be considered as an approximation of the exact visiting probability, we proposed a new estimator based on the idea of approximate importance sampling.
Our estimator is able to overcome the effect of approximate stationary distribution of random walk-based sampling methods on the accuracy of the estimation. We showed both analytically and empirically that the proposed framework is asymptotically unbiased.
\section*{Acknowledgment}
This research has been partially supported by ITRC (Iran Telecommunication Research Center) under grant number 6479/500 (90/4/22).
|
\section{Introduction}
A fundamental question of stellar evolution theory is which stars end
their lives as supernovae. Current theory for isolated massive stars
makes two basic predictions for SNe. For one, the
zero-age-main-sequence mass (M$_{\rm ZAMS}$) and mass-loss history control
whether a SN occurs, and secondly, for single stars there is a clear
mapping between M$_{\rm ZAMS}$\ and the type of SN
\citep{woosley02,heger03a,dessart11}. In particular, the lower masses
explode with their hydrogen envelopes intact (e.g. II-P, II-L, IIn),
and the most massive stars lose much of their envelopes and explode as
hydrogen deficient SNe (e.g. IIb, Ib/c). However, given the
complexity of the underlying physics, especially binary evolution,
winds, and episodic eruptions, it is unclear whether nature obeys the
same well-delineated mass-dependence.
In fact,the relatively high observed rates of H-deficient SNe
\citep{smith11c} and low upper limits on progenitor masses of Type Ibc
SNe \citep{yoon2012,eldridge2013} imply that binary evolution may
figure prominently in producing the H-deficient SNe. Furthermore,
theory predicts that binary evolution can significantly affect the
mapping between initial stellar mass as SNe type
\citep[e.g.,][]{podsiadlowski1992,tutukov1992,nomoto1995,dedonder1998,yoon2010bin,claeys11,dessart11,eldridge2011}.
It is clear that progress in understanding both mass loss and SNe
requires observational constraints linking the progenitor mass with
the eventual SN.
Unfortunately, of the dozens of SNe that have progenitor mass limits,
only 17 have masses measured from a directly detected progenitor (6 of
these 17 overlap this work). \citet{smartt09b} reviewed the mass
distribution of 30 core-collapse SNe progenitors (20 type IIP and 10
others), but only eight of these had measurements beyond an upper
limit. At that point, there were also 4 other nearby SNe progenitors
with full mass constraints
\citep{woosley1988,aldering1994,crockett2008,fraser2010}, bringing the
total to 12. Since 2009, 5 additional SNe have been measured
\citep{maund11,murphy11a,fraser2012,vandyk2012b,maund2013,fraser2013,vandyk2014b},
and some measurements improved \citep{vandyk2012a,maund2014} making
the total 17 measurements.
These mass estimates are based on serendipitous direct imaging of the
progenitor. First, one searches through the HST archive for bright
evolved stars at the SN position. Then, if a star is found, the
endpoints of stellar evolution models which pass through the color and
magnitude of the likely progenitor are used to estimate the star's
initial main sequence mass. If a star is not found, an upper limit on
the progenitor luminosity is measured.
Even with the limited number of measurements available, there is a
hint of a minimum M$_{\rm ZAMS}$\ for explosion and that the least massive
stars explode as SN II-P \citep{smartt09a}. Interestingly, these
measurements also suggested that the maximum mass for SN II-P may be
lower than expected, with some perhaps having been merged binaries
\citep{smartt09a}, although circumstellar dust may also explain the
observations \citep{walmswell2012}. Counter to expectations, some
H-rich SNe (in particular IIn) have been associated with very massive
stars \citep{galyam09,smith11b,smith11c}. While tantalizing, these
initial results are poorly constrained, and even the simple
$\sim$8~M$_{\odot}$ lower mass limit requires more observational
constraints.
While direct imaging of progenitors is the standard method for
progenitor mass estimation, it suffers from a number of limitations.
First among these is the requirement that the precursor imaging
actually exist. The majority of past SNe have neither pre-existing HST
imaging, nor sufficiently accurate astrometry. Consequently, of the
$\sim\!40$ historic SNe within $\sim\!10$ Mpc, only a handful have
identified progenitors. Some future nearby SNe may also lack
precursor imaging due to the limited observations in the HST archive.
The second major limitation is that even when precursor imaging is
available, interpretation of that imaging depends on modeling of the
most uncertain stages of stellar evolution
\citep{gallart05,smartt09a,yoon2010,langer2012,eldridge2013}.
Existing studies estimate the mass of a precursor by fitting endpoints
of stellar evolution models to its color and magnitude; however, an
evolved star's appearance is not well constrained during the final
evolutionary stages. Binarity, mass loss, pulsation, internal mixing,
the formation of dust in stellar winds, and convective instabilities
in shell-burning layers all contribute to systematic and random
uncertainties in such model endpoints. Matching individual endpoints
of stellar evolution models to a single highly-evolved star on the
brink of explosion therefore places weak constraints on the stellar
mass once systematic uncertainties are taken into account.
In this paper, we take a complementary approach that obviates both of
these limitations. We build on a technique developed by several
investigators over the past two decades, starting with measuring ages
for host star clusters
\citep{efremov1991,walborn1993,panagia2000,barth1996,vandyk1999,maiz2004,wang2005,vinko2009,crockett2008},
moving into finding coeval field populations around both SNe and
supernova remnants
\citet[SNRs;][]{badenes2009,gogarten2009,murphy11a,jennings2012}. We
note that, while we have applied this approach to many SNR locations,
historic SNe are less numerous, but more reliable, targets because
they have well-established locations and types, ensuring that every
progenitor mass corresponds to a {\it bona~fide} core-collapse event.
The technique finds the masses of SNe precursors by analyzing the
stellar populations of stars surrounding the SNe. Using
well-established techniques of stellar population modeling, we can
age-date the star formation (SF) event that led to each SN. The
resulting age places strong constraints on the mass of the precursor,
using the well-understood properties of main-sequence stars. As a
result, in cases where the specific stellar population can be
identified, our technique provides a more reliable progenitor age than
direct imaging, as it does not depend on whether the progenitor was a
single star or a binary. Furthermore, the method works even when
there is no imaging prior to the SN, or when the SN position is only
localized to within a few arcseconds. Hence our technique can be
applied to the location of any historic SN that has sufficiently
high-resolution and deep imaging to measure resolved stellar
photometry of the upper main sequence and/or He-burning sequence.
In \S~2, we discuss our sample, the data analyzed in our study, detail
our analysis technique, and demonstrate its efficacy in test cases.
In \S3, we provide our results in the form of age distributions, a
table of masses, and a table of probability distributions for each
progenitor. Finally, \S~4 gives a summary of our findings.
\section{Data}
\subsection{Sample}
We selected all positively-typed historic core-collapse SNe within $8
{\rm{\,Mpc}}$ that also have $\sim$arcsecond accuracy in their positions from
the Asiago Supernova Catalog \citep{barbon1999}; higher positional
accuracy is not needed, due to the large size of the aperture within
which the CMD is constructed.\footnote{One arcsec corresponds to 5 pc
for every Mpc of distance.} We have shown in \citet{murphy11a} that
our method provides results consistent with direct progenitor
detections \citep[as confirmed by][]{vandyk2013} in galaxies as
distant as 8~Mpc. Beyond this limit, our method is not tested, and
therefore we have confined our sample to be within this distance.
We cross-referenced the SNe catalog with the HST archive and
identified SNe that have HST imaging in at least 2 broadband filters
in ACS, WFPC2, or UVIS. Even relatively shallow data can provide
constraints given that, for the most massive $\sim$50~M$_{\odot}$
progenitors, the surrounding populations are likely to have other very
massive stars with M$_{\rm V}{<}{-}$5. We found 22 SNe that match
these requirements. Another SN, 2011dh, has already been analyzed by
our technique in \citet{murphy11a}. Of our sample there are two that
are possibly SN impostors. These are SN1954J \citep{smith2001} and
SN2002kg \citep{weis2005}. We still include these impostors because
the classification as SN impostors are not definitive and it is likely
that these transients are associated with the last stages of stellar
evolution, in which our proposed method to derive progenitor masses is
still of interest. Table~\ref{sample} shows these SNe for which we can
attempt to derive the SFH and progenitor mass, along with the proposal
ID for the dataset used for out photometry.
There were five SNe which we attempted to analyze, but were not able to
constrain with confidence. These had relatively shallow data for
quite distant events (SN~1980K, SN1985F, SN2002ap, SN2002bu, and
SN2003gd) with fewer than 5 stars detected within a 50 pc physical
radius. We do not consider these due to the sparsity of data, although
it is likely that deeper imaging of these locations would yield
photometry that would result in reliable mass estimates. However it is
also possible these progenitors were runaway stars exploding some
distance from their co-eval population \citep[c.f.,][]{eldridge2011}.
\subsection{Analysis Method}
Our method has been described in several other publications, including
its application to SN 2011dh \citep{murphy11a}, 121 SN remnants (SNRs)
in M31 and M33 \citep{jennings2012,zach2}, and an unusual transient in
NGC~300 \citep{gogarten09a}. We provide a description of the method
here as well for convenience.
In brief, we fit SFHs to CMDs of the population surrounding the site
of the SN to determine the age, and thus mass, of the SN progenitor.
The measurements are anchored by the main-sequence stars surrounding
the event; thus, our age estimates do not depend on whether a binary
or single-star progenitor is assumed. Furthermore, the measurements
are not sensitive to any circumstellar dust present around the
progenitor itself.
In the following subsections we first summarize how well the method
works and provide a proof of concept. Then we detail how our
photometry was performed, how the stellar samples were generated, and
how their age distributions were derived.
\subsubsection{Overview}
The method takes advantage of the fact that most stars form in stellar
clusters \citep{lada03} with a common age ($\Delta
t\!\lesssim\!1-4{\rm{\,Myr}}$) and metallicity. Indeed, over 90\% of stars
form in rich clusters containing more than 100 members with
$M\!>\!50$M$_{\odot}$ \citep{lada03}. The stars that formed in a
common event remain spatially correlated on physical scales up to
$\sim\!100{\rm{\,pc}}$ during the $100{\rm{\,Myr}}$ lifetimes of $4$M$_{\odot}$ stars,
even if the cluster is not gravitationally bound \citep{bastian06}; we
have confirmed this expectation empirically in several test cases
\citep{gogarten09a,murphy11a}. Thus, it is reasonable to assume that
most young stars within a 50 parsecs of many SN are coeval. However,
we note that our assumption breaks down for SNe from runaway stars
\citep{eldridge2011}, which would not be coeval with their surrounding
population.
The age of a SN's host stellar population can be recovered from its
color-magnitude diagram (CMD). In the simplest method, one can fit a
single isochrone to an observed CMD and estimate the turnoff mass of
the youngest stars. However, due to the small numbers of massive
stars, one can easily underestimate the mass, since CMDs can show an
apparent turnoff that is fainter than the true turnoff luminosity
simply because of poor Poisson sampling of the upper end of the IMF.
Instead, we adopt more sophisticated methods that take advantage of
the entire CMD. These methods fit superpositions of stellar
populations to reproduce the observed CMD, using the recovery of
artificial stars to generate realistic distributions of stars from
theoretical isochrones. The recovered recent SFH therefore fits not
just the turnoff luminosity, but the full luminosity function of the main
sequence and the blue and red core Helium-burning sequences as well.
Including the well-populated lower end of the main sequence adds
significant statistical weight when interpreting the sparsely sampled
population of massive upper main sequence stars.
The method allows dust extinction to be reliably taken into account.
The main sequence has a well defined color, such that any shift
towards redder colors must be produced by foreground reddening,
allowing the dust extinction to be inferred from the CMD itself.
Differential extinction can be constrained as well, using the observed
widening of the main sequence over what is expected from photometric
errors. The resulting reddening constraints are dominated by the
young stars in which we are most interested.
\subsubsection{Method Validation}
An example of the efficacy of the method is in its application to SN
1987A. We have run our model fits on deep WFPC2 photometry measured
from the archival data of proposal ID 7434 (PI: Kirshner). We fit the
F555W-F814W CMD in the range 12$<$F814W$<$25 as shown in
Figure~\ref{87deep}, and get a well-constrained median
(22.0$^{+2.3}_{-5.8}$ M$_{\odot}$), which is consistent with the mass
(19$\pm 3$ M$_{\odot}$) derived in direct imaging studies
\citep{woosley88}, and with the combined mass of the binary merger
scenario
\citep[16$+$3$\rightarrow$19;][]{podsiadlowski1990,podsiadlowski1992}.
This comparison between techniques provides strong verification of our
proposed method.
While this test is encouraging, the data for 1987A is significantly
deeper than that of our more distant objects. Two more tests,
however, suggest that our technique works even with much shallower
data. First, we ran our model fits on 1987A only including the
photometry for stars brighter than apparent magnitude of 18.5
(absolute magnitude of 0), comparable to the depth of most of our more
distant targets. The resulting median mass was more poorly
constrained (22.8$^{+2.5}_{-14.4}$ M$_{\odot}$), but was still
consistent with the known mass. Thus, we may lose precision with
shallower data, but we can still obtain useful constraints.
In addition to our tests on SN 1987A, we have verified our technique
out to $\sim$8 Mpc by applying it to SN 2011dh in M51
\citep{murphy11a}, for which we found a progenitor mass of
13$^{+2}_{-1}$ M$_{\odot}$. \citet{maund11} identified a progenitor
in archival HST images, which has since vanished \citep{vandyk2013},
and fit the bolometric luminosity to stellar-evolution models and
derived a progenitor mass of 13$\pm 3$ M$_{\odot}$, consistent with
our constraints.
\subsection{Resolved Stellar Photometry}
To generate the CMD, we measure resolved stellar photometry of the
{\it HST} field containing the location of the historic SN. This
photometry was performed using the packages {\tt HSTPHOT} (for WFPC2
data) or {\tt DOLPHOT} \citep[for ACS data;]{dolphin2000}. These
packages perform point spread function fitting optimized for the
undersampled flat-fielded images that come from {\it HST}. All of the
photometry we use has been publicly released to the High-Level Science
Products in the HST archive through the ANGST and ANGRRR programs
(GO-10915 and AR-10945; PI: Dalcanton). The details of the fitting
and culling parameters used are provided in \citet{dalcanton2009} and
the ANGRRR public data archive
\footnote{https://archive.stsci.edu/prepds/angrrr/}. As part of these
programs, hundreds of thousands of artificial star tests were also
performed to assess completeness and photometric accuracy. These
tests consist of inserting a single star into the data, rerunning the
the data reduction, and assessing whether the fake star was recovered,
and if so, how close its measured brightness was to the input
brightness.
\subsubsection{CMD Sample Selection}
To isolate the subset of stars from our catalogs that were co-spatial
with the historic supernovae, we used the coordinates for the SNe from
the Asiago Supernova Catalog \citep{barbon1999}, and galaxy distances
from \citet{dalcanton2009}. We corrected the astrometry in our
catalogs by cross-correlating 2MASS positions for the bright stars in
our catalogs with our positions. Our catalog astrometry is then
corrected such that the star positions agree with those of 2MASS as
precisely as our centroiding for these bright stars will allow
(typically $\sim$0.1$''$). This correction to our catalog astrometry
made sure that our positions were at least as precise as those in the
SNe catalog (within a few tenths of an arcsecond).
With the location and the distance well-measured, we were able to pull
stars measured within a projected radius of 50~pc of each SNe. In the
most distant cases, this radius is only a bit more than 1 arcsecond,
making the necessary precision of astrometry only $\sim$1$''$.
To provide fake star statistics for the photometric completeness and
precision appropriate to our sample, we required a minimum of ten
thousand fake stars into our images. To reach this number, we
included fake stars from a region of up to a factor of 7 larger than
the real stars. In fields where the quality of the data varies
quickly with position, such as near the center of M82, we applied
additional computing resources to obtain more artificial star tests
within the same radius as the stellar sample. However, for almost all
of our fields, changes in stellar density, and therefore photometric
quality, were small over the field, making it possible to use a large
suite of artificial star tests to improve statistics on our CMD
fitting.
\subsubsection{CMD Fitting}
Our CMD fitting process was very similar to that performed in
\citet{jennings2012}. We used the CMD-fitting package MATCH
\citep{dolphin2002,dolphin2013} to fit each CMD with the stellar
evolution models of \citet{girardi2002} with updates in
\citet{marigo2008} and \citet{girardi2010}. The package allows models
to be shifted in temperature and luminosity space to mimic systematic
uncertainties, and it allows differential extinction to be applied to
the models during fitting.
First, we determined the best-fitting amount of differential
extinction to apply when fitting each SN. We fitted the data with a
grid of values for the differential extinction ($dA_{\rm V}$) and
foreground extinction A$_{\rm V}$. We chose the $dA_{\rm V}$ value
that provided the best fit to the data without requiring an $A_{\rm
V}$ value below the known foreground extinction from
\citet{schlegel1998}. We show an example plot summarizing this
extinction determination method in Figure~\ref{dav}.
With the distance, extinction, and differential extinction values
fixed, we fitted the CMD to find the most likely age distribution,
allowing metallicities for the young population in the range of
-0.6$\leq$[Fe/H]$\leq$0.1. Examples for objects with data quality
typical of most of our sample are shown in
Figure~\ref{93J}-\ref{02hh}, where we show an image of our extraction
region, a plot of the CMD of the 50 pc region, and a final cumulative
star formation history from the fitting routine. In
Figure~\ref{correlations1} we plot our final derived masses against
distance and $A_{\rm V}$. The lack of correlations in the derived
masses as a function of these parameters suggests they do not
introduce any significant bias into our measurements.
To assess systematic uncertainties (due to any model deficiencies), we
reran the fitting with several changes to the models, following
\citet{dolphin2012}. We allowed the effective temperature of the
models to vary by $\Delta$log(T$_{eff}$)=0.02. We allow the
bolometric luminosity of the models to vary by
$\Delta$log(L$_{bol}$)=0.17. Furthermore, we allow the differential
extinction applied to the model to vary by ${\Delta}dA_{\rm V}$=0.2 in
cases of high $dA_{\rm V}$ ($>$0.4). We run 100 fits to 100
realizations of the model, varying all of these parameters to account
for systematic uncertainties resulting from our stellar evolution
models and our treatment of the extinction.
Then, to measure the random uncertainties due to number of stars and
depth of the photometry, we use the {\tt hybridMC} task within the
MATCH package as described in \citet{dolphin2013}. This task
determines the star formation rate that would allow an acceptable fit
to the data for each age bin, thus providing robust upper-limits for
bins where the best-fit star formation rates were 0. With both
uncertainty determinations complete, we combine the random and
systematic uncertainties in quadrature using the MATCH routing {\it
zcmerge} to calculate our final uncertainties on the star formation
rates in each age bin.
Next, we use our total uncertainties on the star formation rates in
each age bin to determine the uncertainty on the fraction of stellar
mass present in each age bin back to 50~Myr. We perform 1000 Monte
Carlo realizations of the measured SFH from 50 Myr to the present. We
then calculate the 16\% and 84\% ranges in stellar mass fraction in
each age bin from these tests. We adopt these percentiles as the
uncertainties on fraction of stellar mass formed in each age bin
relative to the total stellar mass produced in the past 50 Myr.
\section{Progenitor Masses}
Once we have the mass fraction (and associated mass fraction
uncertainty) in each age bin, we calculate our first estimate of the
progenitor mass by determining the median age of the best fit. We
then use our uncertainties to determine the age bins consistent with
containing the median to assign uncertainties on that median age.
Finally, we convert these ages to masses by taking the most massive
star remaining in the model isochrone corresponding to the each age
\citep[see][for more details]{jennings2012}. These values provide our
the nominal progenitor mass and associated uncertainties for each SN.
These median masses and associated uncertainties ($\sigma_{med}$) are
provided in Table~\ref{median}.
Although the assignment of a single progenitor age is of interest,
many of our SFHs contain multiple coeval populations, making a more
complex distribution of the mass probability desirable for some
purposes. We therefore have also tabulated the uncertainty for each
progenitor due to the spread in the recovered age distribution
($\sigma_{pop}$). These uncertainties encompass 68\% of the total
population mass (about the median value of the best fit) with ages
$<$50 Myr including uncertainties and account for the full
distribution ages present at the SN location, similar to the technique
adopted in earlier work \citep{murphy11a,jennings2012}. In most cases,
the stellar mass is relatively well confined to a small age range, but
including this second set of uncertainties shows where there are
multiple ages present. For example, the median age of the young
population surrounding SN1994I is well-determined, providing a
high-precision mass measurement of 10.2$\pm$0.7 M$_{\odot}$; however,
there is also a younger population present that represents a
significant fraction of the stellar mass. If the presence of this
population is taken into account, the uncertainties on the progenitor
mass increase substantially to 10.2$^{+59.2}_{-1.8}$. Thus, in this
case, only under the assumption that progenitor was a member of the
dominant young population is the mass of the progenitor
well-constrained. Otherwise, it is only a lower limit.
Looking at these uncertainties, one can determine which SNe would
benefit most from improved photometry data. Large spreads in the 68\%
population mass accompanied by small errors on the median seem to
occur for SNe with few stars in the CMD. For example, SN1951H has
only 11 stars for fitting, a 10\% error on the median, but
$\sigma_{pop}$ values that encompass the full mass range of the
models. Thus, the well-constrained median suggests that the mass
should also be well constrained, but the small number of stars results
in large uncertainties for other age bins which would likely be
reduced with deeper data and a larger number of detected stars.
Finally, to provide detailed probability distributions for all SN, we
tabulate the probability that the progenitor was in each age/mass bin,
given the SFH and associated uncertainties. These probability
distributions are given in Table~\ref{pdfs}, where each mass bin is
assigned a probability which comes from the most likely SFH, and an
associated uncertainty on the probability, which comes from applying
the uncertainties in the SFH to the mass probability distribution.
Thus, in order to account for both sources of uncertainty on the
progenitor mass (the fitting uncertainty and the uncertainty
associated with the intrinsic range of ages), it is necessary to
assign uncertainties to our probabilities. However, in many cases,
the median mass is relatively well defined ($\sigma{<}20$\%), and
provides a simple, though less thorough, constraint on the progenitor
mass.
Six of the events in our sample have previously-measured progenitor
masses from direct imaging (SN1987A, SN1993J, SN2004dj, SN2004et,
SN2005cs, SN2008bk, see Table~\ref{median}), and while our
measurements are less precise in some cases, they are consistent with
the previous measurements in all cases, as shown in Figure~\ref{comp}.
Indeed, in all cases the previous measurements are consistent with our
most optimistic uncertainties---the uncertainty on the median age of
the young population.
\subsection{Extreme Cases}
A few SNe in our samples stand out as extremely challenging of our
ability to measure star formation histories. For example, in
Figure~\ref{04am} we show the image, photometry, and fitting results
for our most heavily extincted location, SN2004am. In this case, even
though there is a very high amount of differential extinction
($dA_{\rm V}$=2.5), the relatively large number of stars provides a
good constraint on the age distribution. Unfortunately, with this
much dust, there is clearly the possibility of a significant number of
more massive stars being completely hidden from the sample, which
would not be accounted for by our method. We cannot account for stars
that are extincted out of our photometry sample. Thus, this amount of
dust may make this result less reliable than many of the others. Such
examples are unlikely to improve without much deeper data to probe to
very high extinctions.
Another extreme case is SN2004et, where we only have 6 stars in the
CMD due to shallow imaging and a far distance (meaning a small
extraction region on the sky). We show our results for this SN in
Figure~\ref{04et}, where the lack of stars results in very large
uncertainties. Although the uncertainties on the progenitor mass are
large, the full range of masses are not allowed by our uncertainties,
which suggest the mass is $>$16 M$_{\odot}$. These uncertainties are
reliable, as the best-mass is well away from (but within the large
errors of) the mass measured from direct imaging. Interestingly, even
with the large uncertainties, the mass constraint is useful since it
rules out masses lower than the best-fit mass from direct imaging.
This example confirms that our uncertainty estimates are reliable, but
also demonstrates that attempting this technique with any fewer stars
is of little value.
Finally, our results in this work add further validation to our
method. For the 6 SNe with we measure here that have literature
measurements, we plot our measurements against those from the
literature in Figure~\ref{comp}. In all cases our measurements are
consistent with previous measurements within the uncertainties, and no
systematic bias is seen.
\subsection{Progenitor Mass Distribution}
We note that our results are consistent with no SN progenitors
$>$20~M$_{\odot}$, as are all of the progenitor mass measurements
currently available in the literature (see references in Section 1).
While we do have some best estimates that are higher mass, their
uncertainties all extend below 20~M$_{\odot}$. Our most massive
central values are for SN2004et and SN1962M, but these only have 75\%
and 82\% probability of being $>$20~M$_{\odot}$. Furthermore, the
direct imaging mass for SN2004et has an upper limit of 20~M$_{\odot}$,
suggesting that the correct mass is indeed at the low end of our
uncertainties. Figure~\ref{mass} plots the masses in ranked order,
along with the expected distribution of masses for a
\citet{salpeter1955} IMF with different upper-mass cutoffs. The large
uncertainties on the high progenitor masses severely limit our ability
to determine the existence of such a cutoff. Thus, our current sample
and data quality does not provide any conclusive evidence that
high-mass stars produce core-collapse supernovae. This lack of
conclusive $>$20~M$_{\odot}$ progenitors is consistent with findings
of several other studies \citep{smartt09a,jennings2012}, hinting that
there could be a ceiling to SN production or a mass range that
under-produces SNe. However, if we can measure a single progenitor
mass $>$20~M$_{\odot}$ with even 20\% precision, constraints on the
progenitor mass distribution would be greatly improved.
\section{Conclusions}
We have constrained the progenitor masses of 17 historic SNe using CMD
fitting of stellar populations measured from HST archival data.
Eleven of these are new constraints, making the total number of
historic SN progenitor masses 28. Even with this dramatic increase in
mass measurements, there is still not a single high-precision
measurement of a progenitor $>$20~M$_{\odot}$, making characterization
of the progenitor mass distribution difficult.
This work represents all that is possible with the current state of
the HST archive. The power of the technique is clear, and we hope
that future studies will be made possible by more and deeper HST
imaging of nearby galaxies containing historic SNe.
Support for this work was provided by NASA through grants AR-13277,
GO-10915, and Hubble Fellowship grant 51273.01 from the Space
Telescope Science Institute, which is operated by the Association of
Universities for Research in Astronomy, Inc., for NASA, under contract
NAS 5-26555. Z.G.J. is supported in part by a National Science
Foundation Graduate Research Fellowship.
|
\section{introduction}
\label{sec:intro}
Neutrino physics has entered the phase of precision measurements. With the upcoming data in the future, the focus is now to determine the missing fundamental parameters such as the neutrino mass hierarchy, the leptonic Dirac CP-violating phase, and the absolute neutrino mass scale. In addition to the standard neutrino parameters, theoretical and phenomenological investigations of beyond-the-Standard-Model scenarios as sub-leading effects have therefore full attention. Such scenarios include non-standard neutrino interactions, unitarity violation, CPT and Lorentz invariance violation, and models with sterile neutrinos. In this work, we will investigate the impact of sterile neutrinos on the fundamental neutrino parameters and how to constrain sterile neutrinos using reactor neutrino experiments such as the ongoing Daya Bay and upcoming JUNO experiments.
Sterile neutrinos, or strictly speaking fermionic SM singlets, are even more elusive than ordinary active neutrinos, since they are supposed to interact through the gravitational force only, and not through the weak force as active neutrinos. However, sterile neutrinos could mix with active neutrinos, which calls for physics beyond the Standard Model. At the moment, there are three experimental results from neutrino oscillation experiments, which give hints that sterile neutrinos could exist. These three results, usually referred to as anomalies, are the LSND (and MiniBooNE) anomaly \cite{Aguilar:2001ty,AguilarArevalo:2008rc,AguilarArevalo:2010wv}, the Gallium anomaly \cite{Giunti:2006bj,Giunti:2010zu,Giunti:2012tn}, and the reactor anomaly \cite{Mention:2011rk}, which all point to sterile neutrinos with mass of the order of 1~eV and small mixing. It should be noted that if such sterile neutrinos exist, they could be produced in the early Universe, and have played an important role in the cosmological evolution. Global fits to data from short-baseline neutrino oscillation experiments suggest that the data can be described by either three active and one sterile (3+1) neutrinos or three active and two sterile (3+2) neutrinos \cite{Kopp:2013vaa}. However significant constraints come from experiments which would appear to disfavor these anomalies \cite{Armbruster:2002mp,Astier:2003gs}. Even in the case in which these anomalies will be explained not by the presence of sterile neutrinos, there are strong indications on the possibility of the existence of sterile neutrinos with masses lower than $1~{\rm eV}$. Moreover, in viewing of the recent detection of B mode polarization from the BICEP2 experiment \cite{Ade:2014xna}, an analysis of the combined CMB data in the framework of LCDM+r models gives $N_{\rm eff} = 4.00 \pm 0.41$ \cite{Giusarma:2014zza}, in favor of the existence of extra radiation. In this work, we will concentrate on the scenario with 3+1 neutrinos analyzing the constraints from reactor neutrino experiments in a wide range of sterile neutrino masses.
From a theoretical point of view, a lot of novel models have been constructed with the aim of embedding sterile neutrinos in a more fundamental framework. Such possibilities include models of extra dimensions with exponentially suppressed sterile neutrino masses, see for instance Ref.~\cite{Kusenko:2010ik,Adulpravitchai:2011rq}. A slightly-breaking flavor symmetry model may generate a neutrino with much smaller mass than the other two, whose masses are allowed by the symmetry. This has been proposed to generate seesaw neutrinos of keV scale in Refs.~\cite{Shaposhnikov:2006nn,Lindner:2010wr}, see also Ref.~\cite{Mohapatra:2005wk}. While the commonly studied flavor models with non-Abelian discrete symmetries cannot easily produce a non-trivial hierarchy between fermion masses, the Froggatt--Nielsen mechanism is capable of such a production~\cite{Froggatt:1978nt}. This has been proposed to generate seesaw neutrinos of eV--keV scales in Refs.~\cite{Barry:2011fp,Merle:2011yv,Barry:2011wb}. Extensions or variants of the canonical type-I seesaw mechanism often contain additional mass scales, which can be arranged to generate light sterile neutrinos~\cite{Barry:2011wb,Zhang:2011vh,Heeck:2012bz,Dev:2012bd}.
The latest measurements of the small mixing angle $\theta_{13}$ have been established by several neutrino oscillation experiments, but it was the Daya Bay experiment that first found a statistical significance of more than 5$\sigma$ confidence level, and therefore, won the hunt for this mixing angle \cite{An:2012eh}. Apart from the standard oscillation picture, the reactor antineutrino experiments could also help us to probe new physics as non-standard effects in neutrino oscillations~\cite{Ohlsson:2008gx,Leitner:2011aa,Adhikari:2012vc,Khan:2013hva,Girardi:2014gna,Bakhti:2014pva}. In this work, we will use the existing data of the Daya Bay experiment as well as the sensitivity of the future JUNO experiment to put constraints on sterile neutrinos using scenarios with 3+1 neutrinos \cite{Palazzo:2013bsa,Bakhti:2013ora,Esmaili:2013yea}.
This work is organized as follows: In Sec.~\ref{sec:neu_osc_prob}, we will analyze neutrino oscillation probabilities with three active neutrinos and one sterile neutrino both (i) analytically and (ii) numerically. Especially, we will consider three cases for the probabilities based on different regimes of the neutrino mass-squared differences, and the effects of sterile neutrinos on the determination of neutrino mass hierarchy. Then, in Sec.~\ref{sec:DB}, we will investigate the impact and signature of one sterile neutrino using the existing data from the Daya Bay experiment \cite{An:2013uza}. Next, in Sec.~\ref{sec:impact_prec}, we will study the sensitivity to sterile neutrino parameters as well as the impact of one sterile neutrino on the precision measurement of the standard neutrino parameters at the JUNO experiment. Finally, in Sec.~\ref{sec:s&c}, we will summarize the results and present our conclusions.
\section{Neutrino Oscillation Probabilities}
\label{sec:neu_osc_prob}
In the presence of $n$ sterile neutrinos, the neutrino mass matrix is an $(n+3) \times (n+3)$ matrix, which can be diagonalized by means of an $(n+3) \times (n+3)$ unitary matrix $U$. In general, one has $(n+3)(n+2)/2$ mixing angles and $(n+2)(n+1)/2$ Dirac phases. In the case of only one sterile neutrino, $U$ is typically parameterized by
\begin{equation}
U = R_{34}\tilde{R}_{24}\tilde{R}_{14}R_{23}\tilde{R}_{13}R_{12}P \,
, \label{eq:UPMNS4x4}
\end{equation}
where the matrix $R_{ij}$ is a rotation by the angle $\theta_{ij}$ in the corresponding $ij$ space, e.g.
\begin{equation}
R_{34} = \begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 &
c_{34} & s_{34} \\ 0 & 0 & -s_{34} & c_{34} \end{pmatrix} \quad {\rm
or} \quad \tilde{R}_{14} = \begin{pmatrix} c_{14} & 0 & 0 &
s_{14}e^{-i\delta_{14}} \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 1 & 0 \\
-s_{14}e^{i\delta_{14}} & 0 & 0 & c_{14} \end{pmatrix}
\end{equation}
with $s_{ij} = \sin\theta_{ij}$ and $c_{ij}=\cos\theta_{ij}$.
The diagonal matrix $P$ contains three Majorana phases, which are irrelevant to our discussion. In this parametrization, one can figure out that
\begin{equation}
\left| U_{e1} \right| = c_{14} c_{13} c_{12} \; , ~~~ \left| U_{e2} \right| = c_{14} c_{13} s_{12} \; , ~~~ \left| U_{e3} \right| = c_{14} s_{13} \; , ~~~ \left| U_{e4} \right| = s_{14} \; ,
\label{eq:Uei}
\end{equation}
indicating that only the mixing angle $\theta_{14}$ enters reactor electron antineutrino oscillations.
The survival probability of electron antineutrinos from nuclear reactors can be written as
\begin{equation}
P_{\bar{e}\bar{e}} \equiv P_{ee} = 1 - 4\sum_{i < j}|U_{e i}|^2 |U_{e j}|^2 \sin^2 \Delta_{ji} \; ,
\label{eq:prob}
\end{equation}
where $\Delta_{ji} \equiv \Delta m^2_{ji} L/(4E)$ denote the oscillation phases, $L$ the baseline length, $E$ the neutrino energy, and $\Delta m^2_{ji} \equiv m^2_j - m^2_i$ the mass-squared difference of two neutrino mass eigenstates $i$ and $j$. Using Eq.~\eqref{eq:Uei}, we can rewrite the survival probability~\eqref{eq:prob} as
\begin{eqnarray}
P_{ee} &=& 1 - c^4_{14} s^2_{12} \sin^2 2\theta_{13} \sin^2 \Delta_{32} - c^4_{14} c^2_{12} \sin^2 2\theta_{13} \sin^2 \Delta_{31} - c^4_{14} c^4_{13} \sin^2 2\theta_{12} \sin^2 \Delta_{21} \nonumber \\
&& - s^2_{13} \sin^2 2\theta_{14} \sin^2 \Delta_{43} - c^2_{13} s^2_{12} \sin^2 2\theta_{14} \sin^2 \Delta_{42} - c^2_{13} c^2_{12} \sin^2 2\theta_{14} \sin^2 \Delta_{41} \; ,
\label{eq:probexp}
\end{eqnarray}
where the oscillation terms are cast into two rows. The first row collects the contributions from active neutrinos, while the second row from sterile neutrinos.
In what follows, we will concentrate on the oscillation probability at the Daya Bay and JUNO setups, in which one of
the standard oscillation modes in the first row of Eq.~\eqref{eq:probexp} dominates the probability.
In addition, the $\Delta_{43}$ mode is further suppressed by both $\theta_{13}$ and $\theta_{14}$, and can therefore be safely neglected.
Thus, the oscillation probability~\eqref{eq:probexp},
in the limit $c^2_{13} = c^2_{14} = 1$, approximates to
\begin{eqnarray}
P_{ee} & \simeq & 1 - s^2_{12} \sin^2 2\theta_{13} \sin^2 \Delta_{32} - c^2_{12} \sin^2 2\theta_{13} \sin^2 \Delta_{31} - \sin^2 2\theta_{12} \sin^2 \Delta_{21} \nonumber \\
&& - s^2_{12} \sin^2 2\theta_{14} \sin^2 \Delta_{42} - c^2_{12} \sin^2 2\theta_{14} \sin^2 \Delta_{41} \; .
\label{eq:probapp}
\end{eqnarray}
\subsection{The Electron Antineutrino Survival Probability at Daya Bay}
Since the baseline length of the Daya Bay detectors is relatively short, the $\Delta_{21}$ related modes are strongly suppressed by $L$, and it is a good approximation to use $\Delta_{32} \simeq \Delta_{31}$ and $\Delta_{42} \simeq \Delta_{41}$. Hence, the oscillation probability~\eqref{eq:probapp} is simplified to
\begin{eqnarray}
P_{ee} \simeq 1 - \sin^2 2\theta_{13} \sin^2 \Delta_{31} - \sin^2 2\theta_{14} \sin^2 \Delta_{41} \; ,
\end{eqnarray}
where terms like $s^2_{13} s^2_{14}$ have been dropped. The last term appears in short-baseline reactor neutrino experiments when $|\Delta m^2_{41}| \gtrsim 10^{-3}~{\rm eV}^2$, and may play an important role in explaining the reactor neutrino anomaly. Furthermore, the sterile neutrino contributions would make significant modifications to the electron antineutrino spectrum. In case of a larger active-sterile mass-squared difference, the second term leads to fast oscillations, which result in a shift of the total observed events.
In the interesting situation that $\Delta m^2_{31} \simeq \Delta m^2_{41}$, the two terms in the oscillation probability can be combined, and one can define an effective mixing angle as $\sin^22\tilde \theta_{13} = \sin^22\theta_{13} +\sin^22\theta_{14}$. In this case, sterile neutrinos induce mimicking effects that add a correction to the observed mixing angle $\theta_{13}$. Accordingly, Daya Bay loses its sensitivity to sterile neutrinos.
\begin{figure}[!t]
\begin{center}
\subfigure{%
\hspace{-1.4cm}
\includegraphics[width=0.64\textwidth]{fig1a.eps} }%
\subfigure{%
\hspace{-2.4cm}
\includegraphics[width=0.64\textwidth]{fig1b.eps} }
\vspace{-1.2cm}
\caption{\it The oscillation probability $P_{ee}$ at Daya Bay as a function of neutrino energy $E$ for $L=2~{\rm km}$ and $\sin^22\theta_{14}=0.1$ (left plot) as well as $\sin^22\theta_{14}=0.01$ (right plot).
SD refers to the standard oscillation probability.}
\label{fig:PeeDB}
\end{center}
\end{figure}
In Fig.~\ref{fig:PeeDB}, we illustrate the oscillation probability at the Daya Bay far detector with baseline length $L=2~{\rm km}$ and mixing $\sin^22\theta_{14}=0.1$ [cf.~Eq.~\eqref{eq:input} for the other standard oscillation parameters]. As one can read off from the plot, in the limit $|\Delta m^2_{41}| \ll |\Delta m^2_{32}|$, the black solid and green dotted curves almost overlap, and hence, Daya Bay has no sensitivity to sterile neutrinos in this mass regime. In the limit $\Delta m^2_{41} \sim \Delta m^2_{31}$, the sterile polluted curve differs from the standard one. However, this difference can be compensated by taking a smaller value for $\theta_{13}$. A combined analysis of reactor and long-baseline experiments are therefore needed to discriminate this ambiguity. In the regime $|\Delta m^2_{41}| \gg |\Delta m^2_{32}|$, the fast oscillations induced by sterile neutrinos lead to a clear distinction to the standard oscillation behavior, and can be well constrained using the current Daya Bay data.
\subsection{The Electron Antineutrino Survival Probability at JUNO}
Different from the Daya Bay setup, the JUNO detector will be located around 50~km away from the nuclear power plant, indicating that the $\Delta_{21}$ oscillation mode is dominating, whereas the $\Delta_{31}$ and $\Delta_{32}$ related oscillation modes become fast oscillations. The sterile neutrino related oscillation modes $\Delta_{4i}$ induce corrections to the neutrino spectrum. Since the JUNO energy resolution is optimized for the determination of the neutrino mass hierarchy, the JUNO detector turns out to be sensitive to mass-squared differences between $10^{-5}~{\rm eV}^2$ and $10^{-2}~{\rm eV}^2$. Above this mass range, the oscillation frequency is too fast to be distinguished, whereas, below this range, the oscillation behavior does not manifest due to the suppression of baseline length and neutrino energy. Therefore, one may consider the following three cases:
\begin{enumerate}
\item The sterile neutrino is nearly degenerate with one of the three active neutrinos, i.e.~$|\Delta m^2_{4i}| < 10^{-5}~{\rm eV}^2 $ (for $i=1,2$, or $3$). The active-sterile mass-squared differences can be ignored in this case, and the $\Delta_{42}$ and $\Delta_{41}$ terms in Eq.~\eqref{eq:probapp} can always be absorbed into the standard oscillation terms. The role of sterile neutrinos is simply to correct the standard neutrino mixing angles, implying loss of sensitivity to sterile neutrinos.
\item In the case of a much larger active-sterile mass-squared difference, i.e.~$|\Delta m^2_{4i}| > 10^{-2}~{\rm eV}^2 $, the fast active-sterile oscillations are actually beyond the resolution limit of JUNO. In this regime, the Daya Bay setup performs a better probe of sterile neutrinos. The reason is that the baseline length of Daya Bay is much shorter than that of JUNO (the Daya Bay baseline is only about 2~\% of the JUNO baseline), and hence, the fast oscillations at Daya Bay is milder, which provides us with a better chance to distinguish the sterile neutrino induced oscillations from the standard ones.
\item In the range $10^{-5}~{\rm eV}^2 < |\Delta m^2_{4i}| < 10^{-2}~{\rm eV}^2$, the observed neutrino spectrum obtains corrections from sterile neutrinos and one would expect a better sensitivity at JUNO.
\end{enumerate}
\begin{figure}[!t]
\begin{center}
\subfigure{%
\hspace{-1.4cm}
\includegraphics[width=0.64\textwidth]{fig2a.eps} }%
\subfigure{%
\hspace{-2.4cm}
\includegraphics[width=0.64\textwidth]{fig2b.eps} }
\vspace{-1.2cm}
\caption{\it The oscillation probability $P_{ee}$ at JUNO as a function of neutrino energy $E$ for $L=52.5~{\rm km}$ and $\sin^22\theta_{14}=0.1$ (left plot) as well as $\sin^22\theta_{14}=0.01$ (right plot). Here the normal mass hierarchy is assumed.
SD refers to the standard oscillation probability.}
\label{fig:PeeJUNO}
\end{center}
\end{figure}
In Fig.~\ref{fig:PeeJUNO}, the sterile neutrino corrections are illustrated for the JUNO setup. When the active-sterile mixing is sizable, the effects of sterile neutrinos become more significant in the large energy regime, in particular for the cases $\Delta m^2_{41} = 10^{-3} ~{\rm eV}^2$ and $\Delta m^2_{41} = 10^{-4} ~{\rm eV}^2$. The shift of the energy spectrum provides us with the possibility to search for sterile neutrinos. For the case of a small value for $\theta_{14}$, the deviation from the standard oscillations is less pronounced, and one needs in principle a challenging experimental setup with a very high precision to detect sterile neutrinos.
Since the major purpose of JUNO is to settle the neutrino mass hierarchy, one may wonder if the presence of sterile neutrinos may affect the determination of the neutrino mass hierarchy at JUNO. To this end, we present the probability difference between the normal and inverted mass hierarchy cases:
\begin{eqnarray}
\Delta P & = & P^{\rm NH}_{ee} - P^{\rm IH}_{ee} \nonumber \\
& \simeq & 2 \sin2\Delta_{21} \left( s^2_{12} \sin^22\theta_{13}\cos\Delta_{31}\sin\Delta_{31} - c^2_{12} \sin^22\theta_{14}\cos\Delta_{42}\sin\Delta_{42} \right) \; ,
\label{eq:deltaP}
\end{eqnarray}
where NH stands for the normal mass hierarchy ($m_3 > m_1$) and IH the inverted mass hierarchy ($m_3 < m_1$). One can clearly observe from Eq.~\eqref{eq:deltaP} that there exists a very interesting situation that in the limit
\begin{eqnarray}
\Delta_{42} & \simeq & \Delta_{31} \; , \label{eq:condition0}\\
s^2_{12} \sin^22\theta_{13} & \simeq & c^2_{12} \sin^22\theta_{14} \; ,
\label{eq:condition}
\end{eqnarray}
the probability difference is equal to zero, i.e.~$\Delta P=0$. In this special case, both normal and inverted mass hierarchy fits would give the same minimal $\chi^2$, and the JUNO setup loses its ability to determine the neutrino mass hierarchy. In other words, if JUNO cannot discriminate between its normal and inverted mass hierarchy analyses, a light sterile neutrino with mass of the order $\Delta m^2_{41} \simeq \Delta m^2_{32}$ and mixing $\sin^22\theta_{14} \simeq 0.04$ could then be the underlying reason.
\begin{figure}[]
\begin{center}
\subfigure{%
\hspace{-1.4cm}
\includegraphics[width=0.64\textwidth]{fig3a.eps} }%
\subfigure{%
\hspace{-2.4cm}
\includegraphics[width=0.64\textwidth]{fig3b.eps} }
\vspace{-1.2cm}
\caption{\it The probability differences $(P^{\rm SD}_{ee})^{\rm NH} - P_{ee}$ (solid curves) and $(P^{\rm SD}_{ee})^{\rm IH} - P_{ee}$ (dotted curves) as functions of neutrino energy $E$ for $\sin^22\theta_{14}=0.04$ and $\Delta m^2_{42} = \Delta m^2_{31}$ (left plot) as well as $\Delta m^2_{42} = 10^{-4} ~{\rm eV}^2$ (right plot), where $P^{\rm SD}_{ee}$ is the standard neutrino oscillation probability. Here the normal mass hierarchy for $P_{ee}$ is assumed.}
\label{fig:PeeHier}
\end{center}
\end{figure}
In Fig.~\ref{fig:PeeHier}, the impact of sterile neutrinos on the mass hierarchy determination is shown. One can observe from the left plot that when the conditions given in Eq.~\eqref{eq:condition} are fulfilled, normal and inverted mass hierarchy fits will give equally good or equally bad fits to experimental data. In contrast, in the general case, the wrong-hierarchy oscillation probability gives a worse fit, which is clearly seen in the right plot of Fig.~\ref{fig:PeeHier}.
\section{Fit to Daya Bay data}
\label{sec:DB}
In this section, we present the relevant features of the Daya Bay experiment and some of the details of our
statistical analysis.
The Daya Bay experimental setup that we take into account consists of six reactors~\cite{An:2013uza}, emitting
antineutrinos $\bar \nu_e$ whose spectra have been recently
estimated in Refs.~\cite{Mueller:2011nm,Huber:2011wv}.
The total flux of arriving $\bar \nu_e$ at the six antineutrino detectors has been estimated
using the convenient parametrization discussed in Ref.~\cite{Mueller:2011nm} and taking into account
all the distances between the detectors and the reactors (summarised in Tab.~2 of Ref.~\cite{An:2013uza}).
For this analysis we use the data set accumulated during 217 days,
which are extracted from Fig.~2 of Ref.~\cite{An:2013zwz}. The antineutrino energy $E$ is reconstructed
by the prompt energy deposited by the positron $E_{ \rm prompt}$ using
the approximated relation \cite{An:2013uza}
$E \simeq E_{\rm prompt} + 0.8 \; {\rm MeV}$.
The energy resolution
function is a Gaussian function, parametrized according to
\begin{eqnarray}
\sigma(E) [\rm MeV]=
\begin{cases}
\gamma \sqrt{E/\rm MeV - 0.8} \, , \; \mbox{for } E > 1.8 \; \rm MeV \, ,\\
\gamma \, , \; \mbox{for } E \leq 1.8 \; \rm MeV\,,\
\end{cases}
\end{eqnarray}
with $\gamma = 0.08$ MeV. The antineutrino cross section for the inverse beta decay process has been taken from Ref.~\cite{Vogel:1999zy}.
The statistical analysis is performed using a modified version of the GLoBES software \cite{Huber:2007ji,Huber:2003pm,Huber:2004ka} and a $\chi^2$ function which takes into account several sources of systematic errors and retrace the one used by the Daya Bay collaboration.
Details can be found in Ref.~\cite{Girardi:2014gna}.
We analyze the sensitivity of the Daya Bay experiment on the sterile parameters and the
effect of $\theta_{14}$ and $\Delta m^2_{41}$ on the determination of $\theta_{13}$ and $\Delta m^2_{31}$. Fit results have been obtained after a marginalization over the parameters that are not shown in the figures.
In particular, we use Gaussian priors defined through the mean value and the 1$\sigma$ error as follows:
\begin{eqnarray}
\sin^2 \theta_{12} & =& 0.306 (1 \pm 5~\%) \; , \nonumber \\
\sin^2 \theta_{13} &=& 0.021 (1 \pm 20~\%) \; , \nonumber \\
\Delta m^2_{21} &=& [7.58 (1 \pm 5~\%)]\times 10^{-5}~{\rm eV^2} \; , \nonumber \\
\left| \Delta m^2_{31} \right| &= & [(2.35 (1 \pm 20~\%)] \times 10^{-3}~{\rm eV^2} \; .
\label{eq:input}
\end{eqnarray}
The central values in Eq.~\eqref{eq:input} have been obtained from Ref.~\cite{Fogli:2011qn}, although with 1$\sigma$ errors slightly
larger to account for possible (unevaluated) effects due to the presence of sterile neutrinos.
The new parameters $\theta_{14}$ and $\Delta m^2_{41}$ are considered as free parameters:
the mass-squared difference is completely unconstrained in the range $(10^{-6},1)$ eV$^2$,
while for the mixing angle we only considered the upper bound
$\theta_{14} < 20^{\circ}$.
In all figures the green dotted-dashed, yellow dotted, and red solid curves refer to 1$\sigma$, 2$\sigma$, and 3$\sigma$ regions in 2 degrees of freedom (dof), respectively.
The results in the $(\sin^2 2\theta_{14},\Delta m^2_{41})$-plane is shown in the left plot of
Fig.~\ref{fig:dmth14th13} after a marginalization over all the standard oscillation
parameters using the priors defined in Eq.~(\ref{eq:input}), in which we can clearly see that at the smallest confidence level
a best fit point emerges at $(\sin^2 2\theta_{14},\Delta m^2_{41}) = (0.012 , 0.039 \, {\rm eV^2})$.
\begin{figure}[]
\begin{center}
\subfigure{%
\hspace{-1.4cm}
\includegraphics[width=0.64\textwidth]{fig4a.eps} }%
\subfigure{%
\hspace{-2.4cm}
\includegraphics[width=0.64\textwidth]{fig4b.eps} }
\vspace{-1.2cm}
\caption{\it Confidence level regions at 1$\sigma$, 2$\sigma$, and 3$\sigma$ for 2 dof,
after performing a fit to the Daya Bay data,
in the $(\sin^2 2\theta_{14},\Delta m^2_{41})$ and $(\sin^2 \theta_{13},\Delta m^2_{41})$-planes presented in the
left and right plots, respectively.}
\label{fig:dmth14th13}
\end{center}
\end{figure}
However, since a relatively large part of the parameter space is still allowed at 2$\sigma$,
it is interesting to analyze the impact of the presence of a third independent mass-squared difference $\Delta m^2_{41}$ on the measurement of $\theta_{13}$.
This is shown
in the right plot of Fig.~\ref{fig:dmth14th13}, obtained after marginalizing over the undisplayed $\theta_{14}$
(limited by $\theta_{14} < 20^{\circ}$) and the other
standard parameters with priors as in Eq.~(\ref{eq:input}).
We can easily recognize the presence of two distinct regions. One for $\Delta m^2_{41}
\lesssim 10^{-3} \, {\rm eV^2}$ and $\Delta m^2_{41}\gtrsim 5 \times 10^{-3} \, {\rm eV^2}$ (at 3$\sigma$)
where, as also outlined in Ref.~\cite{Palazzo:2013bsa}, the measurement of $\theta_{13}$ is quite robust and almost unaffected
by sterile neutrinos. The other for $10^{-3} \,{\rm eV^2}\lesssim \Delta m^2_{41}\lesssim 5 \times 10^{-3} \,{\rm eV^2}$ in which, given the strong
interplay between $\theta_{13}$ and $\theta_{14}$ for $\Delta m^2_{41} \sim \Delta m^2_{31}$
in the oscillation probability, $\theta_{13}$ can also become vanishingly small.
For our purposes, it is enough to study three different cases, shown in Fig.~\ref{fig:th13th14dm41th13}
(obtained marginalizing over the other
standard parameters and on $\theta_{14}$):
$(\sin^2 2\theta_{14},\Delta m^2_{41}) = (10^{-2},10^{-4} \, {\rm eV^2})$ (upper left plot), $(\sin^2 2\theta_{14},\Delta m^2_{41}) = (0.012,0.039 \, {\rm eV^2})$
(upper right plot, corresponding to the best-fit point shown in the left plot of Fig.~\ref{fig:dmth14th13})
and $\Delta m^2_{41} = 2.5\times 10^{-3}$ eV$^2$ with free $\theta_{14}$ (lower plot).
\begin{figure}[]
\begin{center}
\subfigure{%
\hspace{-1.4cm}
\includegraphics[width=0.65\textwidth]{fig5a.eps} }%
\subfigure{%
\hspace{-2.4cm}
\includegraphics[width=0.65\textwidth]{fig5b.eps} }
\subfigure{%
\includegraphics[width=0.65\textwidth]{fig5c.eps} }
\vspace{-0.3cm}
\caption{\it Confidence level regions at 1$\sigma$, 2$\sigma$, and 3$\sigma$ for 2 dof in the
$(\sin^2 \theta_{13},\Delta m^2_{41} )$-plane after performing a fit to the Daya Bay data.
For the left and the right upper plots, the sterile oscillation parameters are fixed to
$(\sin^2 2\theta_{14},\Delta m^2_{41}) = (10^{-2},10^{-4} \, {\rm eV^2})$
and $(\sin^2 2\theta_{14},\Delta m^2_{41}) = (0.012,0.039 \, {\rm eV^2})$, respectively.
The lower plot has been obtained fixing
$\Delta m^2_{41} = 2.5 \times 10^{-3} \, {\rm eV^2}$ and varying freely $\theta_{14}$.
}
\label{fig:th13th14dm41th13}
\end{center}
\end{figure}
As can be observed from the left and the right upper plots of Fig.~\ref{fig:th13th14dm41th13}, the presence of sterile neutrinos does not affect significantly the determination of the standard
oscillation parameters $\theta_{13}$ and $\Delta m^2_{31}$ for mass-squared differences away from the region $10^{-3} \,{\rm eV^2}\lesssim \Delta m^2_{41}\lesssim 5 \times 10^{-3}\, {\rm eV^2}$. On the other hand, for a mass-squared difference within this range we observe in the lower plot a much larger spread of the allowed values of $\theta_{13}$ and $\Delta m^2_{31}$,
as a consequence of $\Delta m^2_{41} \approx \Delta m^2_{31}$.
As we have mentioned below Eq.~(7), the existence of a sterile neutrino could mimick the effects of a large $\theta_{13}$ in this case.
The best-fit $\theta_{13}$ and $\Delta m^2_{31}$ are however in consistent with their true values. Concretely, we have the best-fit values ($\sin^2\theta_{13}$, $\Delta m^2_{31}$)= (0.022, $2.7\times 10^{-3}~{\rm eV}^2$), (0.020, $2.7\times 10^{-3}~{\rm eV}^2$) and (0.021, $2.7\times 10^{-3}~{\rm eV}^2$) for the upper left, upper right and lower plots, respectively.
\section{Sensitivity at JUNO}
\label{sec:impact_prec}
The JUNO experiment~\cite{Li:2013zyd} has been designed to determine the neutrino mass hierarchy, i.e., the sign of $\Delta m^2_{31}$, by observing the disappearance of reactor electron antineutrinos at a distance of $52.5~{\rm km}$. With high statistics of one hundred thousand $\bar{\nu}_e$ events in six years and an excellent energy resolution $\gamma = 0.03~{\rm MeV}$, the JUNO setup will also have a very good sensitivity to the other standard neutrino oscillation parameters, in particular to $\theta_{12}$ and $\Delta m^2_{21}$. In this section, we explore the impact of sterile neutrinos with a mass-squared difference $\Delta m^2_{41}$ ranging from $10^{-6}~{\rm eV}^2$ to $10^{-1}~{\rm eV}^2$ on precision measurements of ($\theta_{12}$, $\Delta m^2_{21}$) and ($\theta_{13}$, $\Delta m^2_{31}$), and the determination of the neutrino mass hierarchy at JUNO. Moreover, the JUNO sensitivity to sterile neutrinos will be studied and compared with the constraint from the Daya Bay data presented in Sec.~\ref{sec:DB}.
Following the approach in Ref.~\cite{Ohlsson:2013nna}, we perform our simulations for the JUNO setup by using the GLoBES software~\cite{Huber:2003pm,Huber:2004ka,Huber:2007ji}. The true values of the relevant standard parameters are taken from the latest global-fit analysis of neutrino oscillation experiments~\cite{Capozzi:2013csa}:
\begin{eqnarray}\label{eq:paras}
\sin^2\theta_{12} & = & 0.308 \pm 0.017 \; , \nonumber \\
\sin^2\theta_{13} & = & 0.0234 \pm 0.002 \; ,\nonumber \\
\Delta m^2_{21} & = & (7.54 \pm 0.26) \times 10^{-5}~{\rm eV}^2 \; , \nonumber \\
\left| \Delta m^2_{31} \right| & = & (2.43 \pm 0.06) \times 10^{-3}~{\rm eV}^2 \; ,
\end{eqnarray}
where $1\sigma$ errors are assumed to be Gaussian and will be incorporated into our simulations as priors for the corresponding parameters. It is worth mentioning that the true values and uncertainties in Eq.~(\ref{eq:paras}) have been obtained by including the Daya Bay data~\cite{Capozzi:2013csa}, in contrast to those in Eq.~(\ref{eq:input}). Since JUNO is very sensitive to $(\theta_{12},\Delta m^2_{21})$, the priors on these are not relevant here. However, the prior knowledge on $(\theta_{13},\Delta m^2_{31})$ from existing reactor neutrino experiments, such as Daya Bay, is important and will be taken into account.
\begin{figure}[!t]
\begin{center}
\subfigure{%
\hspace{-1.4cm}
\includegraphics[width=0.64\textwidth]{fig6a.eps} }%
\subfigure{%
\hspace{-2.4cm}
\includegraphics[width=0.64\textwidth]{fig6b.eps} }
\vspace{-1.2cm}
\caption{\it Illustration for the impact of sterile neutrinos on the experimental sensitivities to $(\sin^2 \theta_{12}, \Delta m^2_{21})$ at JUNO. In our simulations, the true values in Eq.~(\ref{eq:paras}) have been used. The red (dark-gray), orange (gray), and yellow (light-gray) areas stand respectively for the 1$\sigma$, 2$\sigma$, and 3$\sigma$ regions for 2 dof in the case of no sterile neutrinos, while the fit results in the presence of sterile neutrinos are represented by the purple (dotted-dashed), blue (dotted), and cyan (solid) curves. Left plot: For $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.01, 1.0\times 10^{-4}~{\rm eV}^2)$, the best-fit values are $(\sin^2 \theta_{12}, \Delta m^2_{21}) = (0.309, 7.56\times 10^{-5}~{\rm eV}^2)$. Right plot: For $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.012, 3.9\times 10^{-2}~{\rm eV}^2)$, the best-fit values coincide with those in the case of no sterile neutrinos.}
\label{fig:impact_t12_dm21}
\end{center}
\end{figure}
\subsection{The Parameters $\theta_{12}$ and $\Delta m^2_{21}$}
\label{subsec:th12dm12}
In order to illustrate how sterile neutrinos affect the precision measurement of $(\theta_{12}, \Delta m^2_{21})$, we generate neutrino data at JUNO by assuming a light sterile neutrino with $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.01, 1.0\times 10^{-4}~{\rm eV}^2)$. In addition, the true values of the relevant standard parameters are given in Eq.~(\ref{eq:paras}). Then, the generated data are fitted by the standard parameters, with $\theta_{13}$ and $\Delta m^2_{31}$ being marginalized over. As shown in the left plot of Fig.~\ref{fig:impact_t12_dm21}, the best-fit values in this case turn out to be $(\sin^2 \theta_{12}, \Delta m^2_{21}) = (0.309, 7.56\times 10^{-5}~{\rm eV}^2)$ denoted by ``$\times$", which are significantly different from the best-fit values $(\sin^2 \theta_{12},\Delta m^2_{21}) = (0.308, 7.54\times 10^{-5})$ denoted by ``+" in the standard case. The purple dotted-dashed, blue dotted, and cyan solid curves stand for the $1\sigma$, $2\sigma$, and $3\sigma$ contour curves, respectively. The difference between best-fit and true values of $\theta_{12}$ can be well understood from Eq.~(\ref{eq:probapp}), where $\Delta_{42}$ and $\Delta_{41}$ are of the same order of $\Delta_{21}$ and lead to excessive disappearance of reactor antineutrinos. The latter can also be explained by a larger value of $\theta_{12}$, but without sterile neutrinos. On the other hand, for the chosen true values, $|\Delta_{41}| > |\Delta_{21}| > |\Delta_{42}|$ and $\cos^2 \theta_{12} > \sin^2 \theta_{12}$ indicate that sterile neutrinos introduce an additional term of faster oscillations, which can be mimicked by a larger $\Delta m^2_{21}$. However, if the 1$\sigma$ errors of the priors of $\theta_{13}$ and $\Delta m^2_{31}$ are taken of the order of 20\%, the difference between the standard and the nonstandard fits becomes insignificant. For comparison, we present an analysis of JUNO sensitivity in the standard case without sterile neutrinos, and the shaded areas correspond to the $1\sigma$, $2\sigma$, and $3\sigma$ regions, respectively. Given the true values $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.01, 1.0\times 10^{-4}~{\rm eV}^2)$, it is obvious from Fig.~\ref{fig:impact_t12_dm21} that the JUNO sensitivity to $(\theta_{12}, \Delta m^2_{21})$ is essentially not changed, although the best-fit values may deviate from the true values.
If the best-fit values $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.012, 3.9\times 10^{-2}~{\rm eV}^2)$ from Daya Bay data are taken as true values in our simulations, the JUNO sensitivities to $\theta_{12}$ and $\Delta m^2_{21}$ are almost unchanged, as shown in the right plot of Fig.~\ref{fig:impact_t12_dm21}. According to Eq.~(\ref{eq:probapp}), $\Delta m^2_{41} > \Delta m^2_{31} \gg \Delta m^2_{21}$ implies that the contributions from sterile neutrinos can be hidden by the uncertainties of $(\theta_{13}, \Delta m^2_{31})$, in particular for $\sin^2 2\theta_{14} \ll \sin^2 2\theta_{13}$ in our case. For this set of parameters, JUNO is not sensitive enough to place a restrictive constraint.
It is worthwhile to make a comparison between the sensitivity to $(\sin^2 \theta_{12}, \Delta m^2_{21})$ from our simulations and that given by the JUNO Collaboration. In Fig.~\ref{fig:impact_t12_dm21}, the $1\sigma$ error on $\sin^2 \theta_{12}$ is 0.0015 and that on $\Delta m^2_{21}$ is $0.014 \times 10^{-5}~{\rm eV}^2$, corresponding to a precision of $0.49\%$ and $0.19\%$, respectively. In our simulations, only one reactor with thermal power of $35.8~{\rm GW}$ and a flux normalization uncertainty of $3\%$ are considered, and we have ignored the background and other systematics. For the nominal setup and systematic uncertainties considered in Ref.~\cite{Li:2013zyd}, the estimates of the sensitivity to $(\sin^2 \theta_{12}, \Delta m^2_{21})$ from the JUNO Collaboration are $0.54\%$ and $0.24\%$, which are in reasonably good agreement with ours. However, when the bin-to-bin energy uncorrelated uncertainty ($1\%$), the energy linear scale uncertainty ($1\%$), the energy nonlinear uncertainty ($1\%$), and the background ($1\%$) are taken into account, the precisions will be $0.67\%$ and $0.59\%$~\cite{Wang:2014}. Therefore, our simulated sensitivity will be reduced if the background and the above systematic uncertainties are included.
\subsection{The Parameters $\theta_{13}$ and $\Delta m^2_{31}$}
\label{subsec:impact_mass}
\begin{figure}[!t]
\begin{center}
\subfigure{%
\hspace{-1.4cm}
\includegraphics[width=0.64\textwidth]{fig7a.eps} }%
\subfigure{%
\hspace{-2.4cm}
\includegraphics[width=0.64\textwidth]{fig7b.eps} }
\vspace{-1.2cm}
\caption{\it Illustration for the impact of sterile neutrinos on the experimental sensitivities to $(\sin^2 \theta_{13}, \Delta m^2_{31})$ at JUNO. In our simulations, the true values in Eq.~(\ref{eq:paras}) have been used. The red (dark-gray), orange (gray), and yellow (light-gray) areas stand respectively for the 1$\sigma$, 2$\sigma$, and 3$\sigma$ regions for 2 dof in the case of no sterile neutrinos, while the fit results in the presence of sterile neutrinos are denoted by the purple (dotted-dashed), blue (dotted), and cyan (solid) curves. Left plot: For $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.01, 1.0\times 10^{-4}~{\rm eV}^2)$, the best-fit values coincide with those in the case of no sterile neutrinos. Right plot: For $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.012, 3.9\times 10^{-2}~{\rm eV}^2)$, the best-fit values deviate slightly from those in the standard case.}
\label{fig:impact_t13_dm31}
\end{center}
\end{figure}
In a similar way, we now consider the impact of sterile neutrinos on the measurement of $(\theta_{13}, \Delta m^2_{31})$ at JUNO. In Fig.~\ref{fig:impact_t13_dm31}, we show the fit of standard parameters to the data generated by oscillation probabilities in the presence of sterile neutrinos. The fit to the data generated with $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.01, 1.0\times 10^{-4}~{\rm eV}^2)$ is given in the left plot, while that with $(\sin^2 2\theta_{14}, \Delta m^2_{41}) = (0.012, 3.9 \times 10^{-2}~{\rm eV}^2)$ in the right plot.
In the former case, the best-fit value of $\theta_{13}$ in the sterile neutrino case coincides exactly with that in the standard case. Moreover, the $1\sigma$, $2\sigma$, and $3\sigma$ contour curves overlap with the edges of shaded regions, which are obtained by generating neutrino data without sterile neutrinos. The reason is two-fold. First, $\Delta_{41} \approx \Delta_{21} \ll \Delta_{31}$ and the corrections to the standard oscillation probability of three active neutrinos can be absorbed into the uncertainties of $(\theta_{12}, \sin^2 \Delta_{21})$. Second, the JUNO setup itself has limited sensitivity to $(\theta_{13}, \Delta m^2_{31})$.
In the latter case, the deviation from the fit without sterile neutrinos is visible, but insignificant. Due to $\Delta_{41} > \Delta_{31}$, the best-fit point is now shifted to a larger value of $\Delta m^2_{31}$. It is now evident that a light sterile neutrino does not affect the measurement of $(\theta_{13}, \Delta m^2_{31})$ at JUNO, which in any event is not very sensitive to these two parameters.
\subsection{The Neutrino Mass Hierarchy}
\begin{figure}[!t]
\begin{center}
\subfigure{
\hspace{-2cm}
\includegraphics[width=1.2\textwidth]{fig8.eps} }
\vspace{-2cm}
\caption{\it Impact on the determination of neutrino mass hierarchy at JUNO. In our simulations, neutrino data are generated in the NH case. In the upper plot, thick curves refer to the fits with IH, while the corresponding thin curves to those with NH. In the lower plot, the absolute values of differences between the IH and NH fits $\Delta \chi^2_{\rm min} \equiv |\chi^2_{\rm min} ({\rm IH}) - \chi^2_{\rm min} ({\rm NH})|$ have been given for $\Delta m^2_{41} = 1.0 \times 10^{-4}~{\rm eV}^2$ (solid), $5.0\times 10^{-4}~{\rm eV}^2$ (dashed), $2.49\times 10^{-3}~{\rm eV}^2$ (double-dotted), $2.51\times 10^{-3}~{\rm eV}^2$ (dotted-dashed), and $1.0\times 10^{-2}~{\rm eV}^2$ (dotted).}
\label{fig:impact_mass}
\end{center}
\end{figure}
In Fig.~\ref{fig:impact_mass}, we show the JUNO sensitivity to the neutrino mass hierarchy in the presence of sterile neutrinos. In our simulations, the neutrino data are generated in the NH case and the true values of the standard parameters are given in Eq.~(\ref{eq:paras}). Additionally, the true values of $\Delta m^2_{41}$ are specified in the plot, and the black solid, red dashed, blue double-dotted, green dotted-dashed, and brown dotted curves correspond to $\Delta m^2_{41} = 1.0 \times 10^{-4}~{\rm eV}^2$, $5.0\times 10^{-4}~{\rm eV}^2$, $2.49\times 10^{-3}~{\rm eV}^2$, $2.51\times 10^{-3}~{\rm eV}^2$, and $1.0\times 10^{-2}~{\rm eV}^2$, respectively. The fit to the generated neutrino data has been carried out both in the NH and IH cases. In the upper plot, the values $\chi^2_{\rm min}$ of the IH fits are denoted by thick curves, while those of the NH fits by thin curves of the same kind. The absolute values of the differences between the IH and NH fits, namely $\Delta \chi^2_{\rm min} \equiv |\chi^2_{\rm min} ({\rm IH}) - \chi^2_{\rm min} ({\rm NH})|$, are shown in the lower plot. The value of $\Delta \chi^2_{\rm min}$ can be used to measure the capability of the JUNO setup to discriminate between NH and IH.
It is interesting to observe from the lower plot of Fig.~\ref{fig:impact_mass} that $\Delta \chi^2_{\rm min}$ approximately vanishes at $\sin^2 2\theta_{14} = 0.04$ for $\Delta m^2_{41} = 2.51\times 10^{-3}~{\rm eV}^2$, which corresponds to the green dotted-dashed curve. This can be perfectly understood with the help of Eqs.~(\ref{eq:deltaP}) and (\ref{eq:condition}), where one can see that the oscillation probabilities in the NH and IH cases are equal at this point in parameter space. Therefore, JUNO is unable to pin down the neutrino mass hierarchy in this case. Note that there will be another zero point for $\Delta \chi^2_{\rm min}$ around $\sin^2 2\theta_{14} \approx 0.1$. However, now both $\chi^2_{\rm min}$(IH) and $\chi^2_{\rm min}$(NH) are quite large, implying that three active neutrino oscillations in both the NH and IH cases cannot fit the data well. This indicates that the JUNO setup is sensitive enough to constrain or discover a light sterile neutrino with the corresponding mixing parameters. Except for the mass region $\Delta m^2_{41} \approx\Delta m^2_{31}$, sterile neutrinos have little impact on the determination of the neutrino mass hierarchy.
\subsection{The Sensitivity at JUNO}
\begin{figure}[!t]
\begin{center}
\subfigure{
\includegraphics[width=0.8\textwidth]{fig9.eps} }
\vspace{-0.5cm}
\caption{\it Experimental sensitivity to sterile neutrinos at JUNO. The green (dotted-dashed), yellow (dotted), and red (solid) curves correspond to the 1$\sigma$, 2$\sigma$, and 3$\sigma$ contours for 2 dof, respectively. For comparison, the fit to Daya Bay data in Fig.~\ref{fig:dmth14th13} has been reproduced, where the dark (light) shaded area is excluded by Daya Bay at the 3$\sigma$ (2$\sigma$) confidence level.}
\label{fig:junosensitivity}
\end{center}
\end{figure}
Finally, let us proceed to explore the sensitivity of the JUNO setup to the mixing parameters of sterile neutrinos. In our simulations, neutrino data are generated by the standard oscillation probabilities and the true values are given in Eq.~(\ref{eq:paras}). The data are fitted by the general oscillation probability with sterile neutrinos, and all the relevant standard oscillation parameters $(\theta_{12}, \Delta m^2_{21})$ and $(\theta_{13}, \Delta m^2_{31})$ are marginalized over. Our results have been depicted in Fig.~\ref{fig:junosensitivity}, and compared with the fit to the Daya Bay data. The dark (light) shaded area is excluded by Daya Bay at the 3$\sigma$ (2$\sigma$) confidence level. At the $3\sigma$ confidence level, compared to the JUNO setup, the Daya Bay experiment has a better sensitivity to sterile neutrinos with $\Delta m^2_{41} \gtrsim 4.0\times 10^{-3}~{\rm eV}^2$. In the low-mass region, i.e., $\Delta m^2_{41} < 4.0\times 10^{-3}~{\rm eV}^2$, JUNO always dominates over Daya Bay in constraining light sterile neutrinos. In this sense, it is therefore clear that reactor neutrino experiments at short and medium baselines are complementary to each other.
The JUNO setup is most sensitive to the mass region from $\Delta m^2_{41} = 10^{-4}~{\rm eV}^2$ to $\Delta m^2_{41} = 10^{-3}~{\rm eV}^2$, where the limit $\sin^2 2\theta_{14} < 10^{-2}$ can be reached. The sensitivity is significantly diminished for $\Delta m^2_{41} \approx \Delta m^2_{21}$. In this case, the oscillation probability in Eq.~(\ref{eq:probexp}) is reduced to the standard one with two independent neutrino mass-squared differences, where the spectral information is not useful in constraining sterile neutrinos. In the limit of a vanishing $\Delta m^2_{41}$, we obtain $\sin^2 \Delta_{43} \approx \sin^2 \Delta_{31}$ and $\sin^2 \Delta_{42} \approx \sin^2 \Delta_{21}$, implying that the standard neutrino oscillation terms in Eq.~(\ref{eq:probexp}) receive corrections from sterile neutrinos if $\theta_{14}$ is not vanishingly small. Since JUNO has an excellent sensitivity to $\theta_{12}$, it will be able to set an upper bound on $\sin^2 2\theta_{14}$.
It is worthwhile to mention that the experimental constraints on sterile neutrinos exist in the disappearance channel $\bar{\nu}_e \to \bar{\nu}_e$ at reactor neutrino experiments and $\nu_e \to \nu_e$ for solar neutrino experiments. In Ref.~\cite{Kopp:2013vaa}, for $\Delta m^2_{41} \gg 10^{-2}~{\rm eV}^2$, the upper bounds $\sin^2 2\theta_{14} < 0.215$ and $\sin^2 2\theta_{14} < 0.28$ at $95~\%$ confidence level have been derived from long-baseline reactor experiments and from solar plus KamLAND data, respectively. Therefore, our results from the Daya Bay experiment and the future JUNO experiment in Fig.~\ref{fig:junosensitivity} improve the existing bounds in the high-mass region, and provide new constraints in the low-mass region.
\section{Summary and Conclusions}
\label{sec:s&c}
One goal of reactor neutrino experiments is to probe new physics beyond the standard-oscillation paradigm as sub-leading effects in neutrino flavor transitions. Due to high statistical precision and good measurements with the Daya Bay experiment, one can obtain some insight into the hypothesis of sterile neutrinos and put limits on light sterile neutrinos when the active-sterile mass-squared difference is located between $10^{-3}$ and $10^{-1}~{\rm eV}^2$. Restricted by the baseline and energy resolution, the Daya Bay experiment has poor sensitivity to sterile neutrinos with a mass-squared difference below $10^{-3}~{\rm eV}^2$. In contrast, the future JUNO setup features a higher resolution on the neutrino spectrum and has a longer baseline compared to Daya Bay, and hence plays a complementarity role to the current measurements especially in the small mass-squared difference regime. This is particularly relevant for solar neutrinos, since the MSW solution suggests a low energy of the spectra of events at Super-Kamiokande and SNO, which is however not shown in the data. A light sterile neutrino with a mass-squared difference of the order of $10^{-5}~{\rm eV}^2$ and a weak mixing with active neutrinos could explain this suppression \cite{deHolanda:2003tx,deHolanda:2010am}. Furthermore, when the recent detection of B mode polarization from the BICEP2 experiment \cite{Ade:2014xna} is considered, an analysis of the combined CMB data in the framework of LCDM+r models gives $N_{\rm eff}=4.00\pm 0.41$ \cite{Giusarma:2014zza}, which also prefers the existence of extra radiation.
In this work, we have therefore focused on the 3+1 neutrino scenario with only one sterile neutrino and investigated the impact of light sterile neutrinos on short and medium-baseline reactor antineutrino experiments. In particular, we have performed a detail study of antineutrino oscillations and determined the sensitive mass regimes of sterile neutrinos for Daya Bay and JUNO. For both setups, active-sterile neutrino oscillations could in principle mimic the standard oscillations when the active-sterile mass-squared difference is close to one of the standard neutrino mass-squared differences, and hence, one looses sensitivity to sterile neutrinos. Our numerical analysis indicates that the public Daya Bay data suggests an upper limit on the sterile neutrino mixing angle $\sin^22\theta_{14} \lesssim 0.06$ at $3\sigma$ level for the mass-squared difference between $10^{-3}$ and $10^{-1}~{\rm eV}^2$. In addition, for fixed sterile neutrino oscillation parameters, the effects of sterile neutrinos on the determination of $\theta_{13}$ and $\Delta m^2_{31}$ are rather tiny and can be neglected in extracting the standard parameters. Regarding the JUNO setup, the high-energy resolution improves the sensitivity to $\sin^22\theta_{14} \lesssim 0.016$ for $ \Delta m^2_{41} \in (10^{-4} ,10^{-3})~{\rm eV}^2$ and six years of running. However, for a relatively large mass-squared difference, the JUNO sensitivity is not comparable to the one of Daya Bay, due to the longer baseline. When the active-sterile mass-squared difference is around $10^{-4}~{\rm eV}^2$, the measured $\theta_{12}$ and $\Delta m^2_{21}$ deviate from their true values, whereas $\theta_{13}$ and $\Delta m^2_{31}$ are not affected by the sterile neutrino pollution. We have also found a special parameter region that, when $\sin^22\theta_{14} \simeq 0.04$ and $\Delta m^2_{42} \simeq \Delta m^2_{31}$, the sterile neutrino polluted oscillation probability would be almost the same for both NH and IH, indicating that the JUNO setup completely loses its power to discriminate the active neutrino mass hierarchy.
\begin{acknowledgments}
We acknowledge the hospitality and support from the NORDITA scientific program ``News in Neutrino Physics'', April 7--May 2, 2014 during which the initial parts of this study was performed. One of us (I.G.) thanks S. T. Petcov for useful discussions.
This work was supported by the INFN program on ``Astroparticle Physics'' and the European Union FP7-ITN INVISIBLES (Marie Curie Action PITAN-GA-2011-289442-INVISIBLES) (I.G.), MIUR (Italy) under the program Futuro in Ricerca 2010 (RBFR10O36O) (D.M.), the Swedish Research Council (Vetenskapsr{\aa}det), contract no.~621-2011-3985 (T.O.), and the Max Planck Society through the Strategic Innovation Fund in the project MANITOP (H.Z.).
\end{acknowledgments}
|
\section{Introduction. \label{intro}}
Consider the problem
\begin{subnumcases}{\label{linear_problem}}
\Delta f = 0 & in $\Omega$, \label{harmonic_ext_linear} \\
\ddot{f} - \kappa
\overline{\Delta} \partial_\nu f = G & on $\partial \Omega$, \label{eq_f_bry_estimates} \\
f(0, \cdot) = f_0, \, \dot{f}(0, \cdot ) = f_1 & on $\partial \Omega$ \label{ic_f_bry_estimates},
\end{subnumcases}
where $\Omega \subset \mathbb R^n$ is a bounded domain whose boundary $\partial \Omega$
is an
$n-1$-dimensional manifold embedded in $\mathbb R^n$;
$\overline{\Delta}$ is the Laplacian on $\partial \Omega$ and $\Delta$ the Laplacian in
$\mathbb R^n$; $\partial_\nu$ is the outer normal derivative on $\partial \Omega$;
$\kappa$ is a positive constant; $f: [0,T] \times \Omega \rightarrow \mathbb R$ is the
unknown, $T>0$; $G: [0,T] \times \partial \Omega \rightarrow \mathbb R$, $f_0: \partial \Omega \rightarrow \mathbb R$,
and $f_1:\partial \Omega \rightarrow \mathbb R$ are given functions;
and `` $\dot{}$ " means derivative with respect
to $t$, where we write $f = f(t,x)$, $t \in [0,T]$, $x \in \overline{\Omega}$.
We shall elaborate an appropriate notion of weak solution to (\ref{linear_problem}),
then establish the existence and uniqueness of weak solutions on any
time interval $[0,T]$.
Let us discuss some motivations to study (\ref{linear_problem}). In this regard,
it is perhaps worthwhile to start noticing that, from a PDE perspective,
problem (\ref{linear_problem}) is
very natural. Without the normal derivative $\partial_\nu$ on the second
term on the left-hand side of (\ref{eq_f_bry_estimates}), the problem decouples:
(\ref{eq_f_bry_estimates})-(\ref{ic_f_bry_estimates})
becomes a wave equation on the boundary, which can be solved by standard techniques, and
equation (\ref{harmonic_ext_linear}) says that this solution on $\partial \Omega$ is extended to
the interior via the unique harmonic extension of $\left. f \right|_{\partial \Omega}$.
A similar procedure is no longer possible when the term $\partial_\nu$
is present, as in (\ref{linear_problem}). The introduction of the normal derivative
can be viewed as
one of the simplest ways of modifying the wave equation on the boundary as to make
it dependent on the interior values of $f$.
A more direct motivation to investigate (\ref{linear_problem}) is
that it arises at the linearized level in the study of the (incompressible)
free boundary Euler equations, as we now explain.
Consider the motion of an inviscid incompressible fluid
within a bounded region of space, and suppose further that the
boundary of the region confining the fluid is not rigid, being
allowed to move according to the pressure exerted by the fluid
(hence
the name ``free bounary").
This is the situation, for example, in a liquid drop, or in
Newtonian self-gravitating fluid bodies, such as
stars \cite{LindNor, Mak, Nis, White} (the analogue problem for viscous fluids
was first and extensively studied by Solonnikov \cite{MogSol, Sol1, Sol2, Sol3, Sol3,
Sol5, Sol6, Sol7}, with some more recent
advances found in
\cite{KPW, PS, Se1, Se2, Se3} and references therein).
In such situations, the domain containing the fluid changes over time.
One thus
writes $\Omega(t)$, and $\Omega(t)$ becomes one of the unknowns of the problem.
The equations of motion describing the situation of the previous paragraph are
the well-known free boundary Euler equations:
\begin{gather}
\begin{cases}
\frac{\partial u}{\partial t} + \nabla_u u = - \nabla p & \text{ in } \Omega(t), \\
\operatorname{div} (u) = 0 & \text{ in } \Omega(t), \\
p = \kappa \mathcal A & \text{ on } \partial \Omega(t), \\
\langle u, \nu \rangle = v & \text{ on } \partial \Omega(t), \\
u(0) = u_0,
\end{cases}
\label{free_Euler}
\end{gather}
where $u$ is the fluid velocity, $p$ is the fluid pressure,
$\mathcal A$ is the mean curvature
of $\partial \Omega(t)$, $\nu$ is the unit outer normal to $\partial \Omega(t)$,
$v$ is the velocity of the moving boundary $\partial \Omega(t)$, and $\kappa$ a
non-negative constant known as coefficient
of surface tension. We refer the reader to the literature (e.g. \cite{CS, Lin, White})
for a detailed discussion of these equations. It is important to point out that,
despite its importance and the great deal of work dedicated
to (\ref{free_Euler})
\cite{Amborse, AmbroseMasmoudi, ChLin, Craig, E0, Lannes, Lin2, Nalimov, Sch,
ShatahZeng, Wu, Yosihara}, only
recently the problem has been shown to be well-posed \cite{CS, CSB, Lin}
(other recent results, including the study of the compressible free boundary
Euler equations, are \cite{CS2, CS3, CS4, CSH, CSL}).
A very natural question is that of the behavior of solutions to (\ref{free_Euler})
in the limit $\kappa \rightarrow \infty$. Physically, large values of $\kappa$ correspond
to domains with longer relaxation times or, more colloquially, to stiffer domains.
Therefore, one would expect that solutions (\ref{free_Euler}) with large $\kappa$
should be near solutions of the standard Euler equations in the \emph{fixed}
domain $\Omega \equiv \Omega(0)$.
The study of the limit $\kappa \rightarrow \infty$, along with a proof of the corresponding convergence,
was carried out by the author
and David G. Ebin in \cite{DE2d} in the case of two-spatial dimensions (the
reader is also referred to \cite{DE2d} for a more detailed discussion of
the intuition behind this convergence).
The core of the analysis
consists in studying the problem from the point of view of
Lagrangian coordinates, in which case all quantities can be written as time-dependent functions
on the fixed domain $\Omega$ ($=\Omega(0)$). The flow of the vector
field $u$, $\eta(t, \cdot): \Omega \rightarrow \mathbb R^n$,
is decomposed in a part fixing the boundary and a boundary motion. Such
decomposition takes the form
\begin{gather}
\eta = (\operatorname{id} + \nabla f) \circ \beta,
\label{decomp}
\end{gather}
where $\beta$ is a diffeomorphism of the domain $\Omega$ (so in particular $\beta(\partial \Omega) =
\partial \Omega$), $f: \Omega \rightarrow \mathbb R$, and $\operatorname{id}$ is
the identity diffeomorphism. The term $\nabla f$ controls the motion of the boundary, and
using (\ref{free_Euler}), it is possible to derive an equation for $f$.
\emph{At the linearized level and to highest order,
this equation is} (\ref{linear_problem}); the third order operator
$\overline{\Delta} \partial_\nu$ stems from the mean curvature of the
moving boundary. See \cite{DE2d} for details.
In \cite{DE2d}, we were interested in studying the limit $\kappa \rightarrow \infty$,
and therefore we relied on the aforementioned
existence results for (\ref{free_Euler}) (particularly, \cite{CS}). Therefore,
the existence of solutions for the linearized problem, namely, (\ref{linear_problem}), has not been addressed
in \cite{DE2d}.
While it will be shown in a future work that the decomposition (\ref{decomp}) can be employed
to derive existence of solutions to (\ref{free_Euler}) \cite{DE3d}, such an analysis
is based on the calculus of pseudo-differential operators and techniques similar to
\cite{K}. Hence, the simpler, more traditional
methods that we shall present here do not appear
elsewhere. Furthermore, the results in \cite{DE3d} do not cover the case of
\emph{weak} solutions to
(\ref{linear_problem}), which is the main point of this paper (see definition \ref{defi_weak}).
We point out that the singular limit $\kappa \rightarrow \infty$ investigated in
\cite{DE2d,DE3d} fits into the larger picture of properties of solutions viewed
as curves on infinite dimensional manifolds of mappings, which has been extensively
studied in the context of the Euler
equations. See the references \cite{BB, E_manifold, E0, E1, E2, E3, E4, EbinSymp, ED, EM, MEF},
and the discussion in the introduction of \cite{DE2d}. While here we shall not study
the dependence on the parameter $\kappa$, it is instructive to keep the above ideas in mind.
In this regard, compare (\ref{linear_problem}) with the toy-model presented in \cite{E2}.
Naturally, dynamic boundary value problems have a long history,
leading to variants of (\ref{linear_problem}). Adding to the
aforementioned works, whose focus is mainly on equations of hyperbolic type,
the reader can consult, for instance, \cite{Escher, Hintermann}
and references therein, for a point of view that stresses parabolic equations.
Equations involving two time derivatives and a
third order operator also have been studied before (see, for instance,
\cite{Glenn}, and references therein, and see also the related
\cite{TG}). In particular, due to the elliptic operator $\overline{\Delta}$,
the boundary equation
(\ref{eq_f_bry_estimates}) is reminiscent of the so-called Wentzell boundary
conditions, which have been widely studied by
A. Favini, G. Goldstein, J. Goldstein, and S. Romanelli
(a sample of such works is \cite{FGGR1, FGGR2, FGGR3, FGGR4, FGGR5}).
In order to state our results, some notation and definitions are needed.
Fix some $ T > 0$. Define
\begin{gather}
X^3_T(\partial \Omega) = C^0([0,T], H^3(\partial \Omega) )
\cap C^1([0,T], H^{\frac{3}{2}}(\partial \Omega) )
\cap C^2([0,T], H^{0}(\partial \Omega) ),
\nonumber
\end{gather}
and
\begin{gather}
X^\frac{3}{2}_T(\partial \Omega) = C^0([0,T], H^\frac{3}{2}(\partial \Omega) )
\cap C^1([0,T], H^{0}(\partial \Omega) ),
\nonumber
\end{gather}
with norms
\begin{align}
\parallel f \parallel_{X^3_T(\partial \Omega)} & = \sup_{t \in [0,T]} \parallel f \parallel_{3, \partial}
+ \sup_{t \in [0,T]} \parallel \dot{f} \parallel_{\frac{3}{2}, \partial}
+ \sup_{t \in [0,T]} \parallel \ddot{f} \parallel_{0, \partial},
\nonumber
\end{align}
and
\begin{align}
\parallel f \parallel_{X^\frac{3}{2}_T(\partial \Omega)} & = \sup_{t \in [0,T]} \parallel f \parallel_{\frac{3}{2}, \partial}
+ \sup_{t \in [0,T]} \parallel \dot{f} \parallel_{0, \partial}.
\nonumber
\end{align}
Above, $H^s(\partial \Omega)$ is the
Sobolev space whose norm
is denoted by $\parallel \cdot \parallel_{s,\partial}$.
Notice that
\begin{gather}
X_T^3(\partial \Omega) \subset X_T^{\frac{3}{2}}(\partial \Omega).
\nonumber
\end{gather}
The intersections forming $X_T^s(\partial \Omega)$ are of
Sobolev spaces that differ by $\frac{3}{2}$ derivatives. This is because equation
(\ref{eq_f_bry_estimates}) is second order in time and third order in space, thus each
time derivative corresponds to $\frac{3}{2}$ spatial derivatives.
The spaces
$X_T^s(\Omega)$ are similarly defined, and the norm in $H^s(\Omega)$ is denoted $\parallel \cdot \parallel_s$.
As it is implied in the above definitions, we are working with Sobolev spaces
defined with $s \in \mathbb R$. In the case $s \geq 0$, it is useful to have the following
explicit form. Put $s = m + \sigma$, where $m$ is an integer and
$0 < \sigma < 1$. Then
\begin{gather}
\parallel u \parallel_s^2 = \parallel u \parallel^2_m + \left[ D^m u \right]_\sigma^2,
\label{fractional_norm}
\end{gather}
where $\left[ \cdot \right]_\sigma$ is the semi-norm
\begin{gather}
\left[ u \right]_\sigma^2 = \int_{\Omega \times\Omega}
\frac{ |u(x) - u(y) |^2 }{ |x-y|^{n+2\sigma}} \, dx dy,
\nonumber
\end{gather}
with $n$ as the dimension of $\Omega$, and $ \parallel \cdot \parallel^2_m$ as the standard Sobolev
norm defined for integer $m$ \cite{Hitch}. As usual, $H^0$ is simply the $L^2$ space.
As (\ref{linear_problem}) has not appeared in
the literature before, our main
interest is to define a natural notion of weak solutions to problem (\ref{linear_problem}), and
then show that these solutions exist. With this in mind, our treatment
will focus on the simple situation where $\Omega$ is the unit ball,
and we restrict ourselves to the case $n=3$.
In this situation,
problem (\ref{linear_problem}) simplifies considerably, although many
of the arguments below can be extended to a more general setting.
Furthermore, this covers one of
the main cases of interest, namely, that motivated by the linearization of
(\ref{free_Euler}) as discussed above and studied in \cite{DE2d, DE3d}.
We now proceed to state our results.
Denote $L^2(T) = L^2 ( [0,T] \times \partial \Omega )$. Define a map
\begin{align}
\begin{split}
& \mathcal L: X^3_T(\partial \Omega) \rightarrow L^2(T) \times H^\frac{3}{2}(\partial \Omega) \times H^0(\partial \Omega)
\equiv \mathcal H \\
& \mathcal L(f) = (\ddot{f} - \kappa \overline{\Delta} \partial_\nu f, f(0), \dot{f}(0)),
\end{split}
\nonumber
\end{align}
where $\partial_\nu f$ is computed using
the harmonic extension of $f$ to $\Omega$,
and we write $f(0) = f(0,\cdot)$, $\dot{f}(0) = \dot{f}(0,\cdot)$. Let $\mathcal R \subset \mathcal H$
be the image of $\mathcal L$. We shall prove the following.
\begin{prop}
Let $\Omega \subset \mathbb R^3$ be the ball of radius one centered at the origin,
and fix some $T>0$.
Let $\mathcal L$, $\mathcal R$, and $\mathcal H$ be as above. Then:\\
\noindent (i) $\mathcal L$ is injective, and since
$X^3_T(\partial \Omega) \subset X^\frac{3}{2}_T(\partial \Omega)$,
$\mathcal L^{-1}$ defines a map
\begin{gather}
\mathcal L^{-1}: \mathcal R \rightarrow X^\frac{3}{2}_T(\partial \Omega).
\nonumber
\end{gather}
The map $\mathcal L^{-1}$ is continuous as a map from $\mathcal R$ to $X^\frac{3}{2}_T(\partial \Omega)$. \\
\noindent (ii) The closure of $\mathcal R$ in $\mathcal H$, denoted $\overline{\mathcal R}$,
is the whole of $\mathcal H$, i.e.,
$\overline{\mathcal R} = \mathcal H$, and $\mathcal L^{-1}$ extends to a continuous
linear map, $\overline{\mathcal L^{-1}}$, from $\mathcal H$ to $X^\frac{3}{2}_T(\partial \Omega)$. \\
\noindent (iii) The image of $\overline{\mathcal L^{-1}}$ is
\begin{gather}
\overline{X^3_T(\partial \Omega)\,}{}^{X^\frac{3}{2}_T(\partial \Omega)},
\nonumber
\end{gather}
i.e., the closure of $X^3_T(\partial \Omega) \subset X^\frac{3}{2}_T(\partial \Omega)$
in the $X^\frac{3}{2}_T(\partial \Omega)$ topology.
\label{prop_extension}
\end{prop}
We can now introduce the following.
\begin{defi}
Let $f_0 \in H^\frac{3}{2}(\partial \Omega)$, $f_1 \in H^0(\partial \Omega)$, and $G \in L^2(T)$ be given.
We say that
\begin{gather}
f \in \overline{X^3_T(\partial \Omega)\,}{}^{X^\frac{3}{2}_T(\partial \Omega)} \bigcap
X^2_T( \Omega)
\nonumber
\end{gather}
is a weak solution
of (\ref{linear_problem}), if $\left. f \right|_{\partial \Omega} = \overline{\mathcal L^{-1}}\left( (G, f_0, f_1 )\right)$,
where $\overline{\mathcal L^{-1}}: \mathcal H \rightarrow \overline{X^3_T(\partial \Omega)\,}{}^{X^\frac{3}{2}_T(\partial \Omega)}$
is the map given by proposition \ref{prop_extension}, and
$f$ satisfies (\ref{harmonic_ext_linear}) in $\Omega$.
\label{defi_weak}
\end{defi}
To understand why this is a suitable definition of weak solutions for problem (\ref{linear_problem}),
one should think of the example
of the wave equation. In that case, given initial data in $H^1(\mathbb R^n) \times H^{0}(\mathbb R^n)$ and an inhomogeneous
term in, say, $L^1([0,T], H^{0}(\mathbb R^n))$, the weak solution $u$ is in
$C^0([0,T],H^1(\mathbb R^n)) \cap C^1([0,T], H^{0}(\mathbb R^n))$. Thus, the weak solution
has one less spatial derivative than the order of the equation, with $\partial_t u$
one degree less differentiable in space than $u$ itself. Such a regularity is
a consequence of the energy estimate, in which an integration by parts is performed.
In our case, each time derivative corresponds to $\frac{3}{2}$ spatial ones, and
we heuristically think of integrating by parts half of the derivatives of the
third order spatial term.
Proposition \ref{prop_extension} essentially contains the existence of weak solutions,
but we state it separately for convenience.
\begin{theorem}
Let $\Omega \subset \mathbb R^3$ be the ball of radius one centered at the origin,
and fix some $T>0$. Given $G \in L^2(T)$, $f_0 \in H^\frac{3}{2}(\partial \Omega)$,
and $f_1 \in H^0(\partial \Omega)$, there exists a unique weak solution to the
problem (\ref{linear_problem}).
\label{main_theorem}
\end{theorem}
It should be stressed that for sufficiently regular data, existence for
(\ref{linear_problem}) can probably be derived by other means.
The novelty
of theorem \ref{main_theorem} is centered around the notion of weak
solutions and their existence. In this regard, it is important to stress
that a semi-group approach can also be employed to study (\ref{linear_problem}),
in which case one is led, via Stone's theorem, to investigate the existence
of mild-solutions to the problem \cite{ABHN, Yosida}. Such mild solutions are,
in fact, closely
related to our notion of weak solution. We believe, however, that the energy-method
approach here employed is significantly simpler in the sense that it does not rely
on heavy functional-analytic techniques, and is also of independent interest
to the community more acquainted with such type of estimates.
\section{Energy estimates.}
In this section we carry out the necessary energy estimates for the proofs
of proposition \ref{prop_extension}
and theorem \ref{main_theorem}. We start recalling some useful tools and fixing some of
the notation.
\subsection{Auxiliary results.\label{auxiliary}}
Here, we collect some well known facts that will be used in
the paper. Their proofs can be found in many sources, e.g.,
\cite{Adams, BB, Hitch, E1, P}.
First, recall that restriction to the boundary gives rise to
a bounded linear map,
\begin{gather}
\parallel u \parallel_{s, \partial} \leq C \parallel u \parallel_{s + \frac{1}{2}},~~ s > 0,
\label{restriction}
\end{gather}
with $C =C(n,s,\Omega)$.
The usual interpolation inequality will also be needed: if $s_1 < s_2 < s_3$, then
\begin{gather}
\parallel u \parallel_{s_2} \leq \parallel u \parallel_{s_1}^\frac{s_3 - s_2}{s_3-s_1} \parallel u \parallel_{s_3}^\frac{s_2-s_1}{s_3-s_1}.
\label{interpolation}
\end{gather}
We finally recall the standard Cauchy inequality with $\gamma$,
\begin{gather}
ab \leq \gamma a^2 + \frac{1}{4\gamma} b^2,
\label{Cauchy_epsilon}
\end{gather}
$\gamma > 0$ (this inequality is usually called Cauchy inequality with
$\varepsilon$,
with the letter $\varepsilon$ used instead of $\gamma$. We shall reserve
$\varepsilon$ for other purposes below, thus we use $\gamma$ in (\ref{Cauchy_epsilon}) to avoid confusion).
\subsection{Coordinates and notation.}
Here, we make some remarks about coordinates and notation.
Recalling that $\Omega$ is the ball of radius one centered at the origin, we write
\begin{gather}
\partial \Omega_r = \partial B_r(0).
\nonumber
\end{gather}
Sometimes, we employ spherical coordinates $(r,\phi,\theta)$, so that
\begin{gather}
\Delta = \partial_r^2 + \frac{2}{r} \partial_r + \frac{1}{r^2} \Delta_{S^2},
\label{Laplacian_spherical}
\end{gather}
where $\Delta_{S^2}$ is the Laplacian on the standard round sphere, given
in these coordinates
by
\begin{gather}
\Delta_{S^2} = \partial_\phi^2 + \frac{\cos \phi}{\sin \phi} \partial_\phi
+ \frac{1}{\sin^2 \phi} \partial^2_\theta.
\nonumber
\end{gather}
In particular,
the Laplacian on $\partial \Omega_r$, which we denote $\overline{\Delta}$ for any $r$, is
\begin{gather}
\overline{\Delta} = \frac{1}{r^2} \Delta_{S^2} .
\label{induced_Laplacian_spherical}
\end{gather}
We shall use $\overline{\Delta}$ as an operator on the whole of $\Omega$.
To be precise, this is not defined
at zero, but the origin can be removed without changing the value of the integrals
$\int_\Omega$ containing $\overline{\Delta}$ that will appear below.
In particular, $\overline{\Delta} f$ is defined
on $\Omega$ (but the origin).
We shall also make use of the following coordinate choice.
For $\varepsilon > 0$, let
\begin{gather}
\Omega_\varepsilon = \Omega \backslash (B_\varepsilon(0) \cup C_\varepsilon ),
\nonumber
\end{gather}
where $C_\varepsilon$ is the cone given in spherical coordinates by
$\{ \phi \geq \pi - \varepsilon \}$.
Choose Fermi coordinates $\{ x^\mu \}_{\mu=1}^3$ at the north
pole of $\partial \Omega$. These coordinates
cover $\Omega_\varepsilon$, and the Euclidean metric takes the form
$g = (g_{\alpha \beta})$, with
$g_{33} = 1$, $g_{i3} = 0$, $i=1,2$, and $g_{ij}$, $i,j=1,2$,
being the metric induced on
the
level sets $\{ x^3 = \text{ constant} \}$, which in turn correspond
to $\partial \Omega_r \cap \Omega_\varepsilon$.
Furthermore, $\partial_3$ is orthogonal to $\partial \Omega_r \cap \Omega_\varepsilon$,
and $\partial_3 = - \partial_r$. We illustrate the construction of these coordinates
in figure \ref{figure_Om_e}, where we also depict further notation
that will be used below.
\begin{figure}[!ht]
\centering
\rotatebox{0.0}{\includegraphics[scale=.45]{Om_e.eps}}
\caption{Illustration of the set $\Omega_\varepsilon$ and its
boundary $\partial \Omega_\varepsilon = \partial \Omega_\varepsilon^1 \cup
\partial \Omega_\varepsilon^2$.}
\label{figure_Om_e}
\end{figure}
For the rest of the paper, the following convention is adopted.
\begin{notation}
Greek indices run from $1$ to $3$ and Latin indices from $1$ to $2$. The letter
$C$ will be used to denote several different constants, as usual.
\end{notation}
In the above coordinates, equation (\ref{harmonic_ext_linear}) then reads
\begin{gather}
\nabla^\mu \nabla_\mu f = 0,
\nonumber
\end{gather}
where $\nabla$ is covariant differentiation in the Euclidean metric, but written in
this system of coordinates. Notice that all covariant derivatives will commute, as
the metric is flat, and we shall use this in the calculations below.
\subsection{Basic energy inequality.\label{section_basic_energy}}
For the rest of this section, let $f \in X^3_T(\partial \Omega)$
be a solution to (\ref{eq_f_bry_estimates})-(\ref{ic_f_bry_estimates}).
We also denote by $f$ its harmonic extension to $\Omega$, i.e.,
$f$ satisfies (\ref{harmonic_ext_linear}) in $\Omega$.
Define the energy
\begin{gather}
E = \frac{1}{2} \int_{\partial \Omega} \left( \dot{f}^2
- \kappa f\overline{\Delta} \partial_\nu f \right).
\label{def_E}
\end{gather}
Differentiating
and integrating by parts,
\begin{gather}
\dot{E} = \int_{\partial \Omega} \dot{f} \ddot{f} -
\frac{1}{2}\kappa \int_{\partial \Omega} \dot{f} \overline{\Delta} \partial_\nu f
-
\frac{1}{2}\kappa \int_{\partial \Omega} \overline{\Delta} f \partial_\nu \dot{f}.
\label{E_1}
\end{gather}
Let $\varphi = \varphi(r)$ be a sufficient regular function such that
$\varphi(1) = 1$, i.e., $\varphi \equiv 1$ on $\partial \Omega$.
Apply
Green's identity
\begin{gather}
\int_\Omega \left( u \Delta v - v \Delta u \right) = \int_{\partial \Omega}
\left( u \frac{\partial v }{\partial \nu} - v \frac{\partial u}{\partial \nu} \right)
\nonumber
\end{gather}
with $v = \varphi \overline{\Delta} f$ (which equals $\overline{\Delta} f$ on $\partial \Omega$)
and $u = \dot{f}$, to get
\begin{align}
\begin{split}
\int_{\partial \Omega} \overline{\Delta } f \partial_\nu \dot{f}
=
\int_{\partial \Omega} \varphi \overline{\Delta } f \partial_\nu \dot{f}
& =
\int_{\partial \Omega} \dot{f} \overline{\Delta } \partial_\nu f +
\int_{\partial \Omega} \dot{f} [\partial_\nu, \varphi \overline{\Delta} ] f
-
\int_\Omega \dot{f} \Delta
(\varphi \overline{\Delta} f ) ,
\end{split}
\label{E_2}
\end{align}
where $[\cdot, \cdot]$ is the commutator, we have used that (\ref{harmonic_ext_linear}) implies $\Delta \dot{f} = 0$, and that $\varphi = 1$ on $\partial \Omega$.
Using (\ref{E_2}) into (\ref{E_1}) gives
\begin{gather}
\dot{E} = \int_{\partial \Omega} \dot{f} \left( \ddot{f} -
\kappa \overline{\Delta} \partial_\nu f \right )
+ r_1
\label{dot_E}
\end{gather}
where
\begin{gather}
r_1 = -\frac{1}{2} \kappa
\int_{\partial \Omega} \dot{f} [\partial_\nu, \varphi \overline{\Delta} ] f
+ \frac{1}{2} \kappa
\int_\Omega \dot{f} \Delta
(\varphi \overline{\Delta} f ) .
\nonumber
\end{gather}
To analyze $r_1$, pick $\varphi(r) = r^2$, which satisfies the previous assumptions on
$\varphi$. Then, on $\partial \Omega$,
\begin{gather}
\partial_\nu (\varphi\overline{\Delta} f ) = \partial_r(r^2 \overline{\Delta} f )
= \partial_r \Delta_{S^2} f = \Delta_{S^2} \partial_r f =
\overline{\Delta} \partial_\nu f = \varphi \overline{\Delta} \partial_\nu f,
\nonumber
\end{gather}
which implies $[\partial_\nu, \varphi \overline{\Delta} ] = 0$.
Using also (\ref{Laplacian_spherical}),
\begin{align}
\begin{split}
\Delta (\varphi \overline{\Delta} f) & = \Delta \Delta_{S^2} f
= \Delta_{S^2} \partial_r^2 f
+ \Delta_{S^2}\left( \frac{2}{r} \partial_r f \right)
+ \Delta_{S^2} \left (\frac{1}{r^2} \Delta_{S^2} f \right)
= \Delta_{S^2} \Delta f = 0.
\end{split}
\nonumber
\end{align}
Thus $r_1 = 0$ and (\ref{dot_E}) becomes
\begin{gather}
\dot{E} = \int_{\partial \Omega} \dot{f} \left( \ddot{f} -
\kappa \overline{\Delta} \partial_\nu f \right ).
\label{dot_E_2}
\end{gather}
Invoking (\ref{eq_f_bry_estimates}) and the Cauchy-Schwarz inequality,
(\ref{dot_E_2}) gives
\begin{gather}
\dot{E}
\leq \frac{1}{2} \parallel \dot{f} \parallel_{0,\partial}^2 + \frac{1}{2} \parallel G \parallel_{0,\partial}^2,
\label{dot_E_CS}
\end{gather}
where we recall that $\parallel \cdot \parallel_{s,\partial}$ is the Sobolev
norm on the boundary.
\begin{lemma}
\begin{gather}
-\int_{\partial \Omega} \overline{\Delta} f \partial_\nu f \geq 0.
\nonumber
\end{gather}
\label{positivity_lemma}
\end{lemma}
\begin{proof}
Integrating by parts,
\begin{gather}
-\int_{\partial \Omega} \overline{\Delta} f \partial_\nu f
= \int_{\partial \Omega} \langle \nabla_\partial f, \nabla_\partial \partial_\nu f \rangle,
\label{first_by_parts_positivity}
\end{gather}
where $\nabla_\partial$ is the gradient on $\partial \Omega$.
We need to commute $\nabla_\partial$ and
$\partial_\nu$.
This is easily done in spherical coordinates, yielding
\begin{gather}
\left[ \nabla_\partial, \nabla_\nu \right] = \frac{1}{r} \nabla_\partial .
\nonumber
\end{gather}
(\ref{first_by_parts_positivity}) becomes
\begin{align}
\begin{split}
-\int_{\partial \Omega} \overline{\Delta} f \partial_\nu f
& =
\int_{\partial \Omega} \frac{1}{r} |\nabla_\partial f|^2
+
\int_{\partial \Omega} \langle \nabla_\partial f, \nabla_\nu \nabla_\partial f \rangle .
\end{split}
\label{second_by_parts_positivity}
\end{align}
Choose Fermi coordinates as explained at the beginning
of this section. Then $\nabla^\mu \nabla_\mu f = 0$
implies $\nabla^\mu \nabla_\mu \nabla_\sigma f = 0$,
and $\nabla^\sigma f \nabla^\mu \nabla_\mu \nabla_\sigma f =0$.
Integrating over $\Omega_\varepsilon$ and integrating by parts,
\begin{align}
\begin{split}
0 = \int_{\Omega_\varepsilon}
\nabla^\sigma f \nabla^\mu \nabla_\mu \nabla_\sigma f
& = - \int_{\Omega_\varepsilon} \nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
+ \int_{\partial \Omega_\varepsilon} \nabla^\sigma f \nabla_V \nabla_\sigma f ,
\end{split}
\nonumber
\end{align}
so
\begin{align}
\begin{split}
\int_{\Omega_\varepsilon} \nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
& = \int_{\partial \Omega_\varepsilon^1} \nabla^\sigma f \nabla_V \nabla_\sigma f
+ \int_{\partial \Omega_\varepsilon^2} \nabla^\sigma f \nabla_V \nabla_\sigma f,
\end{split}
\label{split_partial_1_2}
\end{align}
where $\partial \Omega_\varepsilon^1 =
\partial \Omega \backslash (\partial \Omega_\varepsilon \cap \partial \Omega)$,
$\partial \Omega_\varepsilon^2 = \partial \Omega_\varepsilon \backslash \partial \Omega_\varepsilon^1$
(see figure \ref{figure_Om_e}),
and $\nabla_V$ is covariant differentiation in the normal direction, with $V$ the unit
outer normal. On $\partial \Omega_\varepsilon^1$, $\nabla_V = -\nabla_3$. Compute
\begin{align}
\begin{split}
\nabla_3 \nabla_\sigma f & =
\partial_3 \partial_\sigma f - \Gamma_{3\sigma}^j \partial_j f,
\end{split}
\nonumber
\end{align}
where we used that any Christoffel symbol with two indices $3$ vanishes in
Fermi coordinates. Let $\overline{\nabla}$ be the covariant derivative on
$\partial \Omega$. Then, since $\nabla_i f = \partial_i f = \overline{\nabla}_i f$,
\begin{align}
\begin{split}
\nabla_3 \nabla_i f & =
\nabla_3 \overline{\nabla}_i f.
\end{split}
\nonumber
\end{align}
Hence,
\begin{align}
\begin{split}
\int_{\partial \Omega_\varepsilon^1} \nabla^\sigma f \nabla_V \nabla_\sigma f
& =
- \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_3 \overline{\nabla}_i f
- \int_{\partial \Omega_\varepsilon^1} \nabla^3 f \nabla_3 \nabla_3 f \\
\end{split}
\label{cov_diff_split_i_n}
\end{align}
Using (\ref{cov_diff_split_i_n}) into (\ref{split_partial_1_2}) gives
\begin{align}
\begin{split}
\int_{\Omega_\varepsilon} \nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
& =
- \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_3 \overline{\nabla}_i f
- \int_{\partial \Omega_\varepsilon^1} \nabla^3 f \nabla_3 \nabla_3 f
+ \int_{\partial \Omega_\varepsilon^2} \nabla^\sigma f \nabla_V \nabla_\sigma f,
\end{split}
\label{intermediate_partial_1}
\end{align}
Taking $\nabla_3$ of $\nabla^\mu \nabla_\mu f =0$, commuting the
covariant derivatives, multiplying by $\nabla^3 f$ and
integrating by parts,
\begin{align}
\begin{split}
0 = \int_{\Omega_\varepsilon}
\nabla^3 f \nabla^\mu \nabla_\mu \nabla_3 f
& = - \int_{\Omega_\varepsilon} \nabla^\mu \nabla^3 f \nabla_\mu \nabla_3 f
+ \int_{\partial \Omega_\varepsilon} \nabla^3 f \nabla_V \nabla_3 f ,
\end{split}
\nonumber
\end{align}
so, since $\nabla_V = - \nabla_3$ on $\partial \Omega_\varepsilon^1$,
\begin{align}
\begin{split}
\int_{\partial \Omega_\varepsilon^1} \nabla^3 f \nabla_3 \nabla_3 f
& =
-\int_{\Omega_\varepsilon} \nabla^\mu \nabla^3 f \nabla_\mu \nabla_3 f
+ \int_{\partial \Omega_\varepsilon^2} \nabla^3 f \nabla_V \nabla_3 f
\end{split}
\nonumber
\end{align}
Using the above expression for
$\int_{\partial \Omega_\varepsilon^1} \nabla^3 f \nabla_3 \nabla_3 f$ into
(\ref{intermediate_partial_1}),
\begin{align}
\begin{split}
\int_{\Omega_\varepsilon} \nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
& =
- \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_3 \overline{\nabla}_i f
+\int_{\Omega_\varepsilon} \nabla^\mu \nabla^3 f \nabla_\mu \nabla_3 f
\\
& - \int_{\partial \Omega_\varepsilon^2} \nabla^3 f \nabla_V \nabla_3 f
+ \int_{\partial \Omega_\varepsilon^2} \nabla^\sigma f \nabla_V \nabla_\sigma f,
\end{split}
\nonumber
\end{align}
so that
\begin{align}
\begin{split}
\int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_V \overline{\nabla}_i f
=
- \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_3 \overline{\nabla}_i f
& =
\int_{\Omega_\varepsilon} \nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
- \int_{\Omega_\varepsilon} \nabla^\mu \nabla^3 f \nabla_\mu \nabla_3 f
\\
& + \int_{\partial \Omega_\varepsilon^2} \nabla^3 f \nabla_V \nabla_3 f
- \int_{\partial \Omega_\varepsilon^2} \nabla^\sigma f \nabla_V \nabla_\sigma f.
\end{split}
\label{estimate_bry_interior_epsilon}
\end{align}
The first two terms on the right hand side combine to give
\begin{gather}
\int_{\Omega_\varepsilon} \left( \nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
- \nabla^\mu \nabla^3 f \nabla_\mu \nabla_3 f \right)
= \int_{\Omega_\varepsilon} \left( |\nabla\nabla f|^2 - |\nabla \nabla_V f |^2\right) \geq 0.
\nonumber
\end{gather}
For the integrals along $\partial \Omega_\varepsilon^2$, write
$\partial \Omega_\varepsilon^2 = \partial \Omega_\varepsilon^{21} \cup \partial \Omega_\varepsilon^{22}$,
where
\begin{gather}
\partial \Omega_\varepsilon^{21} = \partial \Omega_\varepsilon^{2} \cap \partial B_\varepsilon(0),
\nonumber
\end{gather}
and
\begin{gather}
\partial \Omega_\varepsilon^{22} = \partial \Omega_\varepsilon^{2} \backslash
\partial \Omega_\varepsilon^{21}.
\nonumber
\end{gather}
Then
\begin{align}
\begin{split}
\left|
\int_{\partial \Omega_\varepsilon^{21}} \nabla^3 f \nabla_V \nabla_3 f
- \int_{\partial \Omega_\varepsilon^{21}} \nabla^\sigma f \nabla_V \nabla_\sigma f
\right| \leq &
\int_{\partial \Omega_\varepsilon^{21}} \left| \nabla^3 f \nabla_V \nabla_3 f \right|
+ \int_{\partial \Omega_\varepsilon^{21}} \left| \nabla^\sigma f \nabla_V \nabla_\sigma f \right|
\\
\leq &
\int_{\partial B_\varepsilon(0)} \left| \nabla^3 f \nabla_V \nabla_3 f \right|
+ \int_{\partial B_\varepsilon(0)} \left| \nabla^\sigma f \nabla_V \nabla_\sigma f \right| \\
\leq &
C \int_{\partial B_\varepsilon(0)} \parallel f \parallel^2_{2,\partial B_\varepsilon(0)} \\
\leq & C \int_{\partial B_\varepsilon(0)} \parallel f \parallel^2_{\frac{5}{2}} \\
\leq & C \parallel f \parallel^2_{\frac{5}{2}} \varepsilon^2,
\end{split}
\label{error_ball}
\end{align}
where $\parallel \cdot \parallel_{2,\partial B_\varepsilon(0)}$ is the Sobolev norm on $\partial B_\varepsilon(0)$,
and on the next-to-the-last step we used (\ref{restriction}).
Similarly,
\begin{align}
\begin{split}
\left|
\int_{\partial \Omega_\varepsilon^{22}} \nabla^3 f \nabla_V \nabla_3 f
- \int_{\partial \Omega_\varepsilon^{22}} \nabla^\sigma f \nabla_V \nabla_\sigma f
\right| \leq &
C \parallel f \parallel^2_{\frac{5}{2}} \varepsilon.
\end{split}
\label{error_cone}
\end{align}
From (\ref{error_ball}) and (\ref{error_cone}) we see that the integrals over $\partial \Omega_\varepsilon^2$
vanish in the limit $\varepsilon \rightarrow 0^+$. Hence,
\begin{gather}
\int_{\partial \Omega} \langle \nabla_\partial f, \nabla_\nu \nabla_\partial f \rangle
= \lim_{\varepsilon \rightarrow 0^+} \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_V \overline{\nabla}_i f
\geq 0,
\label{positivity_limit_epsilon}
\end{gather}
where we recall that $\nu$ is the outer unit normal to $\partial \Omega$.
Combining (\ref{positivity_limit_epsilon})
with (\ref{second_by_parts_positivity}) yields the result.
\end{proof}
As a consequence of lemma \ref{positivity_lemma}, we have
$E(t) \geq 0$, and
\begin{gather}
\parallel \dot{f} \parallel_{0, \partial}^2 \leq E.
\label{f_dot_partial_E}
\end{gather}
Thus, (\ref{dot_E_CS}) gives
\begin{gather}
\dot{E} \leq \frac{1}{2} \parallel G \parallel_{0,\partial}^2 + \frac{1}{2} E,
\nonumber
\end{gather}
or,
\begin{gather}
E(t) \leq E(0) + \frac{1}{2} \int_0^t \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau
+ \frac{1}{2} \int_0^t E(\tau) \, d\tau,
\nonumber
\end{gather}
which gives, after iteration,
\begin{gather}
E(t) \leq \left( E(0) + \frac{1}{2} \int_0^t \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau
\right) e^{\frac{1}{2} t}.
\label{simple_energy_estimate_linear}
\end{gather}
From
(\ref{harmonic_ext_linear}) and standard elliptic theory we have
\begin{gather}
\parallel f \parallel_s \leq C \parallel f \parallel_{s-\frac{1}{2},\partial}.
\label{elliptic_estimate_Laplace_eq}
\end{gather}
And invoking elliptic theory once more, (\ref{elliptic_estimate_Laplace_eq}) then gives
the following estimate:
\begin{align}
\begin{split}
\left| \int_{\partial \Omega} f \overline{\Delta} \partial_\nu f \right |
& = \left| \int_{\partial \Omega} \overline{\Delta}f \partial_\nu f \right | \\
& \leq \parallel \overline{\Delta}f \parallel_{-\frac{1}{2},\partial} \parallel \partial_\nu f \parallel_{\frac{1}{2},\partial}
\\
& \leq C\parallel f \parallel_{\frac{3}{2}, \partial } \parallel f \parallel_2 \\
& \leq C \parallel f \parallel_{\frac{3}{2}, \partial}^2.
\end{split}
\label{energy_less_3_2}
\end{align}
Thus, (\ref{energy_less_3_2}) and the definition of $E$, i.e. (\ref{def_E}), imply
\begin{gather}
E \leq C \parallel f \parallel_{\frac{3}{2}, \partial }^2 + C \parallel \dot{f} \parallel_{0,\partial}^2.
\label{energy_less_norm}
\end{gather}
Combining (\ref{energy_less_norm}) at time zero with (\ref{simple_energy_estimate_linear})
produces
\begin{gather}
E(t) \leq C \left(
\parallel f(0) \parallel_{\frac{3}{2}, \partial }^2 + C \parallel \dot{f}(0) \parallel_{0,\partial}^2
+ \frac{1}{2} \int_0^t \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau
\right) e^{\frac{1}{2} t}.
\label{simple_energy_estimate_linear_2}
\end{gather}
Next, we show that
\begin{lemma}
\begin{gather}
-\int_{\partial \Omega} f \overline{\Delta} \partial_\nu f =
- \int_{\partial \Omega}
\overline{\Delta} f \partial_\nu f \geq C \parallel f \parallel^2_{\frac{3}{2}, \partial }
-C \parallel f \parallel_{0,\partial}^2.
\nonumber
\end{gather}
\label{lemma_3_2_norm}
\end{lemma}
\begin{rema}
The presence of the negative term $-\parallel f \parallel_{0,\partial}^2$ is necessary as
$\int_{\partial \Omega} \overline{\Delta} f \partial_\nu f$ is zero on constant functions.
\end{rema}
\begin{proof}
The equality follows by integration by parts. In light of
(\ref{second_by_parts_positivity}), it suffices to obtain the inequality for
\begin{align}
\begin{split}
\int_{\partial \Omega} \langle \nabla_\partial f, \nabla_\nu \nabla_\partial f \rangle ,
\end{split}
\nonumber
\end{align}
which, as in lemma
\ref{positivity_lemma},
along $\partial \Omega_\varepsilon^1$ corresponds to the term
\begin{gather}
- \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_3 \overline{\nabla}_i f.
\nonumber
\end{gather}
Proceed as in lemma \ref{positivity_lemma}
until (\ref{estimate_bry_interior_epsilon}), and consider its
first two terms on the right hand side. Their integrand gives
\begin{gather}
\nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
- \nabla^\mu \nabla^3 f \nabla_\mu \nabla_3 f
= \nabla^\mu \nabla^i f \nabla_\mu \nabla_i f.
\nonumber
\end{gather}
Letting $\overline{\nabla}$ denote the covariant derivative along
$\partial \Omega_r \cap \Omega_\varepsilon$ (which correspond to the level sets
$x^3 = \text{constant}$), noticing that $\overline{\nabla}_i f = \partial_i f
= \nabla_i f$ and that, therefore, $\overline{\nabla}_i$ gives a well-defined
operator on $\Omega_\varepsilon$, we can write the above as
\begin{gather}
\nabla^\mu \nabla^\sigma f \nabla_\mu \nabla_\sigma f
- \nabla^\mu \nabla^3 f \nabla_\mu \nabla_3 f
= | \nabla \overline{\nabla} f|^2,
\nonumber
\end{gather}
thus
(\ref{estimate_bry_interior_epsilon}) gives,
\begin{align}
\begin{split}
- \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_3 \overline{\nabla}_i f
& =
\int_{\Omega_\varepsilon} |\nabla \overline{\nabla} f|^2
+ \int_{\partial \Omega_\varepsilon^2} \nabla^3 f \nabla_V \nabla_3 f
- \int_{\partial \Omega_\varepsilon^2} \nabla^\sigma f \nabla_V \nabla_\sigma f.
\end{split}
\label{initial_integral_lemma_3_2}
\end{align}
But
\begin{align}
\begin{split}
\int_{\Omega_\varepsilon} |\nabla \overline{\nabla} f|^2
\geq
\parallel \overline{\nabla} f \parallel_1^2 - C\parallel f\parallel_1^2
\geq
\parallel \overline{\nabla} f \parallel_1^2 - C\parallel f\parallel_{\frac{1}{2}, \partial_\varepsilon}^2
\geq
C \parallel \overline{\nabla} f \parallel_{\frac{1}{2},\partial_\varepsilon}^2
- C\parallel f\parallel_{\frac{1}{2}, \partial_\varepsilon}^2 ,
\end{split}
\label{second_der_f_lemma_3_2}
\end{align}
where in the last step we used (\ref{restriction}), in the next-to-the-last,
(\ref{elliptic_estimate_Laplace_eq}), and
$\parallel \cdot \parallel_{s, \partial_\varepsilon}$ is the Sobolev norm on the
boundary $\partial \Omega_\varepsilon$. Applying the interpolation inequality
(\ref{interpolation}) with $s_1 = 0$, $s_2 = \frac{1}{2}$, and $s_3 = 1$,
\begin{align}
\begin{split}
\parallel f\parallel_{\frac{1}{2}, \partial_\varepsilon}^2 \leq &
\frac{C}{\gamma} \parallel f \parallel_{0,\partial_\varepsilon}^2
+ \gamma \parallel f \parallel_{1,\partial_\varepsilon}^2 ,
\end{split}
\label{interpolation_lemma_3_2}
\end{align}
where the Cauchy inequality with $\gamma$, (\ref{Cauchy_epsilon}),
has been employed.
Using (\ref{interpolation_lemma_3_2}) in (\ref{second_der_f_lemma_3_2}) and recalling
(\ref{fractional_norm}),
\begin{align}
\begin{split}
\int_{\Omega_\varepsilon} |\nabla \overline{\nabla} f|^2
\geq &
C \parallel \overline{\nabla} f \parallel_{\frac{1}{2},\partial_\varepsilon}^2
- C \gamma \parallel f \parallel_{1,\partial_\varepsilon}^2
- \frac{C}{\gamma} \parallel f \parallel_{0,\partial_\varepsilon}^2 \\
= &
C \left[ \overline{\nabla} f \right]_{\frac{1}{2}, \partial_\varepsilon}^2 +
C \parallel f \parallel_{1,\partial_\varepsilon}^2
- C \gamma \parallel f \parallel_{1,\partial_\varepsilon}^2
- \frac{C}{\gamma} \parallel f \parallel_{0,\partial_\varepsilon}^2 \\
\geq &
C \left[ \overline{\nabla} f \right]_{\frac{1}{2}, \partial_\varepsilon}^2 +
C \parallel f \parallel_{1,\partial_\varepsilon}^2
- \frac{C}{\gamma} \parallel f \parallel_{0,\partial_\varepsilon}^2,
\end{split}
\label{second_der_f_lemma_3_2_2}
\end{align}
where the last step follows by choosing $\gamma$ sufficiently small.
Since $\overline{\nabla}$ is differentiation along $\partial \Omega_\varepsilon$,
recalling (\ref{fractional_norm}) once more, we see that
\begin{gather}
C \left[ \overline{\nabla} f \right]_{\frac{1}{2}, \partial_\varepsilon}^2 +
C \parallel f \parallel_{1,\partial_\varepsilon}^2
\geq C \parallel f \parallel_{\frac{3}{2}, \partial_\varepsilon}^2,
\nonumber
\end{gather}
so that (\ref{second_der_f_lemma_3_2_2}) gives
\begin{align}
\begin{split}
\int_{\Omega_\varepsilon} |\nabla \overline{\nabla} f|^2
\geq & C\parallel f \parallel^2_{\frac{3}{2}, \partial_\varepsilon}
- C \parallel f \parallel^2_{0,\partial_\varepsilon},
\end{split}
\nonumber
\end{align}
and therefore (\ref{initial_integral_lemma_3_2}) implies
\begin{align}
\begin{split}
- \int_{\partial \Omega_\varepsilon^1} \overline{\nabla}^i f \nabla_3 \overline{\nabla}_i f
& \geq
C\parallel f \parallel^2_{\frac{3}{2}, \partial_\varepsilon}
- C \parallel f \parallel^2_{0,\partial_\varepsilon}
+ \int_{\partial \Omega_\varepsilon^2} \nabla^3 f \nabla_V \nabla_3 f
- \int_{\partial \Omega_\varepsilon^2} \nabla^\sigma f \nabla_V \nabla_\sigma f.
\end{split}
\label{last_step_lemma_3_2}
\end{align}
To finish the proof, split the first two terms on the right hand
of (\ref{last_step_lemma_3_2}) in integrals along $\partial \Omega_\varepsilon^{21}$ and
$\partial \Omega_\varepsilon^{22}$. Arguing as in lemma \ref{positivity_lemma},
all integrals on $\partial \Omega_\varepsilon^2$ vanish in the limit $\varepsilon\rightarrow 0^+$,
which gives the result.
\end{proof}
As a consequence of lemma \ref{lemma_3_2_norm} and the definition
\ref{def_E}, we have
\begin{align}
\begin{split}
\parallel f \parallel^2_{\frac{3}{2},\partial} & \leq \frac{C}{\kappa} E
+ C \parallel f \parallel_{0,\partial}^2.
\end{split}
\label{estimate_f_3_2_Energy_1}
\end{align}
Using the fundamental theorem of calculus and the Cauchy-Schwarz inequality,
\begin{align}
\begin{split}
\parallel f \parallel_{0,\partial}^2 & \leq
C \parallel f(0) \parallel_{0,\partial}^2 + C \left( \int_0^t \parallel \dot{f} \parallel_{0,\partial} \right)^2
\leq C \parallel f(0) \parallel_{0,\partial}^2 + C t \int_0^t \parallel \dot{f} \parallel_{0, \partial}^2.
\end{split}
\label{estimate_f_3_2_Energy_2}
\end{align}
In the above, we used Jensen's inequality,
\begin{gather}
h\left( \Xint- f \right) \leq \Xint- h(f),
\nonumber
\end{gather}
where $h$ is a convex function and $\Xint-$ the average over the domain of integration,
to estimate
\begin{align}
\begin{split}
\left( \int_0^t \parallel \dot{f} \parallel_{0,\partial} \right)^2
& =
t^2 \left( \Xint-_0^t \parallel \dot{f} \parallel_{0,\partial} \right)^2
\leq t \int_0^t \parallel \dot{f} \parallel_{0, \partial}^2.
\end{split}
\nonumber
\end{align}
Thus, (\ref{f_dot_partial_E}),
(\ref{estimate_f_3_2_Energy_1}), and (\ref{estimate_f_3_2_Energy_2}) give
\begin{align}
\begin{split}
\parallel f \parallel^2_{\frac{3}{2},\partial} & \leq \frac{C}{\kappa} E
+ C \parallel f(0) \parallel_{0,\partial}^2 + C t \int_0^t E.
\end{split}
\label{estimate_f_3_2_Energy_bound}
\end{align}
Inequalities (\ref{f_dot_partial_E}) and (\ref{estimate_f_3_2_Energy_bound})
combined with (\ref{simple_energy_estimate_linear_2}) give bounds
for $\parallel \dot{f} \parallel_{0,\partial}$ and $\parallel f \parallel^2_{\frac{3}{2},\partial}$.
More precisely, we have
\begin{gather}
\parallel \dot{f} \parallel_{0, \partial}^2 \leq C \left(
\parallel f(0) \parallel_{\frac{3}{2}, \partial }^2 + \parallel \dot{f}(0) \parallel_{0,\partial}^2
+ \frac{1}{2} \int_0^t \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau
\right) e^{\frac{1}{2} t},
\label{bound_f_dot}
\end{gather}
and
\begin{align}
\begin{split}
\parallel f \parallel^2_{\frac{3}{2},\partial} & \leq
C \parallel f(0) \parallel_{0,\partial}^2 + \frac{C}{\kappa}
\left(
\parallel f(0) \parallel_{\frac{3}{2}, \partial }^2 + \parallel \dot{f}(0) \parallel_{0,\partial}^2
+ \frac{1}{2} \int_0^t \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau
\right) e^{\frac{1}{2} t} \\
& + C t^2 e^{\frac{1}{2}T} \left(
\parallel f(0) \parallel_{\frac{3}{2}, \partial }^2 + \parallel \dot{f}(0) \parallel_{0,\partial}^2
+ \int_0^t \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau
\right),
\end{split}
\label{bound_f_3_2_}
\end{align}
where we used $e^{\frac{1}{2} t} \leq e^{\frac{1}{2} T}$, and
\begin{gather}
\int_0^t \int_0^\tau \parallel G(\sigma) \parallel_{0,\partial}^2 \, d\sigma d\tau
\leq
\int_0^t \int_0^t \parallel G(\sigma) \parallel_{0,\partial}^2 \, d\sigma d\tau
= t \int_0^t \parallel G(\sigma) \parallel_{0,\partial}^2 \, d\sigma.
\nonumber
\end{gather}
\section{Proofs.}
We are now ready to proof proposition \ref{prop_extension} and
theorem \ref{main_theorem}.
Let $X^{3,+}_T(\partial \Omega)$ be the subspace of $X^3_T(\partial \Omega)$
consisting of functions
such that $f(0) = \dot{f}(0) = 0$, and
$X^{3,-}_T(\partial \Omega)$ be the subspace of $X^3_T(\partial \Omega)$ consisting of functions
such that $f(T) = \dot{f}(T) = 0$.
Recall that
$X_T^3(\partial \Omega) \subset X_T^{\frac{3}{2}}(\partial \Omega)$.
\begin{lemma}
$X^{3,\pm}_T(\partial \Omega)$ is dense in $L^2(T)$.
\end{lemma}
\begin{proof}
This is very similar to the proof that compactly supported functions are dense
in Sobolev spaces, using mollifiers.
\end{proof}
\noindent \emph{Proof of proposition \ref{prop_extension}-(i):}
From
(\ref{bound_f_dot}) and (\ref{bound_f_3_2_}) with $G = 0$, $f(0) = 0$, and $\dot{f}(0)=0$,
it follows that $\mathcal L$ is injective.
Let $(G, f_0, f_1) \in \mathcal H$ and $f = \mathcal L^{-1}\big( (G, f_0, f_1) \big)$. Then
(\ref{bound_f_3_2_}) gives
\begin{gather}
\parallel f \parallel_{\frac{3}{2}, \partial} \leq C(T,\kappa)
\left(
\parallel f_0 \parallel_{\frac{3}{2}, \partial}
+\parallel f_1 \parallel_{0,\partial} + \sqrt{ \int_0^T \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau }
\right),
\nonumber
\end{gather}
where $C(T, \kappa)$ is a constant that depends on $T$ and $\kappa$, and
we have used that
\begin{gather}
\int_0^t \parallel G(\tau) \parallel_{0,\partial}^2 \ d\tau
\leq
\int_0^T \parallel G(\tau) \parallel_{0,\partial}^2 \ d\tau, \, t \in [0,T],
\nonumber
\end{gather}
and that the exponential is an increasing function.
But
\begin{gather}
\int_0^T \parallel G(\tau) \parallel_{0,\partial}^2 \, d\tau
= \int_0^T \int_\Omega |G(\tau,x)|^2 \, dx\, d\tau \leq C \parallel G \parallel_{L^2(T)}^2,
\nonumber
\end{gather}
so that
\begin{gather}
\parallel f \parallel_{\frac{3}{2}, \partial} \leq C(T,\kappa)
\left(
\parallel f_0 \parallel_{\frac{3}{2}, \partial}
+\parallel f_1 \parallel_{0,\partial} + \parallel G \parallel_{L^2(T)}
\right).
\nonumber
\end{gather}
Similarly, from (\ref{bound_f_dot}) we also conclude,
\begin{gather}
\parallel \dot{f} \parallel_{0, \partial} \leq C(T,\kappa)
\left(
\parallel f_0 \parallel_{\frac{3}{2}, \partial}
+\parallel f_1 \parallel_{0,\partial} + \parallel G \parallel_{L^2(T)}
\right).
\nonumber
\end{gather}
These last two inequalities imply
\begin{gather}
\parallel f \parallel_{X^\frac{3}{2}_T(\partial \Omega)} \leq C(T,\kappa)
\left(
\parallel f_0 \parallel_{\frac{3}{2}, \partial}
+\parallel f_1 \parallel_{0,\partial} + \parallel G \parallel_{L^2(T)}
\right).
\label{continuity_L_inverse}
\end{gather}
As the right hand side of this last inequality is the norm of
$(G,f_0,f_1)$ in the topology of $\mathcal H$, we conclude that $\mathcal L^{-1}$ is a continuous map.
\qed \\
\noindent \emph{Proof of proposition \ref{prop_extension}-(ii):}
Denoting by $\overline{\mathcal R}$ the closure
of $\mathcal R$ in $\mathcal H$, $\mathcal L^{-1}$ extends, by continuity, to a continuous linear map
\begin{gather}
\overline{\mathcal L^{-1}}: \overline{\mathcal R} \rightarrow X^\frac{3}{2}_T(\partial \Omega).
\nonumber
\end{gather}
In order to show that the closure of $\mathcal R$ in $\mathcal H$ is the whole of $\mathcal H$, i.e.,
\begin{gather}
\overline{\mathcal R} = \mathcal H,
\nonumber
\end{gather}
we will prove that if $v \perp \overline{\mathcal R}$, then
$ v = 0$. As $\mathcal R$ is dense in its closure, it suffices to show that
\begin{gather}
\text{if } (v,w)_\mathcal H = 0 \text{ for all } w = (G,f_0, f_1) \in \mathcal R,
\text{ then } v = 0,
\nonumber
\end{gather}
where $(\cdot, \cdot, \cdot)_\mathcal H$ is the inner product
on $\mathcal H \equiv L^2(T) \times H^\frac{3}{2}(\partial \Omega) \times
H^0(\partial \Omega)$. Write $v = (H, v_0, v_1)$, $w = w = (G,f_0, f_1)$, and suppose that
\begin{gather}
(H,G)_{L^2(T)} + (v_0, f_0)_{\frac{3}{2}, \partial} + (v_1, f_1)_{0,\partial} = 0
\text{ for all } (G,f_0,f_1) \in \mathcal R,
\nonumber
\end{gather}
where $(\cdot, \cdot)_{L^2(T)}$, $(\cdot, \cdot)_{\frac{3}{2},\partial}$,
and $(\cdot, \cdot)_{0,\partial}$ are the inner products in
$ L^2(T)$, $H^\frac{3}{2}(\partial \Omega)$, and $H^0(\partial \Omega)$, respectively.
By definition of $\mathcal R$ and the fact that $\mathcal L$ is injective, the above means
\begin{gather}
(H,\ddot{f} - \kappa \overline{\Delta} \partial_\nu f)_{L^2(T)}
+ (v_0, f(0))_{\frac{3}{2}, \partial} + (v_1, \dot{f}(0))_{0,\partial} = 0
\text{ for all } f \in X^3_T(\partial \Omega).
\label{inner_product_zero_1}
\end{gather}
Assume first that $f \in X^{3,+}_T(\partial \Omega)$. In this case
(\ref{inner_product_zero_1}) becomes
\begin{gather}
(H, \ddot{f} - \kappa \overline{\Delta} \partial_\nu f)_{L^2(T)} = 0.
\label{inner_product_zero_2}
\end{gather}
Suppose, further, that $H \in X^{3,-}_T(\partial \Omega)$. Then
\begin{align}
\begin{split}
(H, \ddot{f} - \kappa \overline{\Delta} \partial_\nu f)_{L^2(T)} & =
\int_{[0,T]\times \partial \Omega} H ( \ddot{f} - \kappa \overline{\Delta} \partial_\nu f )
\\
& =
\int_{ \partial \Omega} \int_0^T H \ddot{f} \, dt dx -
\kappa \int_{ \partial \Omega} \int_0^T H \overline{\Delta} \partial_\nu f \, dt dx
\end{split}
\label{integration_parts_L2_T}
\end{align}
Integrating by parts in the time variable the first term,
\begin{align}
\begin{split}
\int_{ \partial \Omega} \int_0^T H \ddot{f} \, dt dx
& =
\int_{ \partial \Omega} \int_0^T \ddot{H} f \, dt dx,
\end{split}
\label{integration_parts_time_H_f}
\end{align}
where we used that $f \in X^{3,+}_T(\partial \Omega)$ and
$H \in X^{3,-}_T(\partial \Omega)$.
For the second term in (\ref{integration_parts_L2_T}),
switch the order of integration and integrate by parts the Laplacian term to get
\begin{gather}
\int_{ \partial \Omega} \int_0^T H \overline{\Delta} \partial_\nu f \, dt dx
=
\int_0^T \int_{ \partial \Omega} \overline{\Delta} H \partial_\nu f \, dx dt.
\nonumber
\end{gather}
Next, consider the harmonic extensions of $f$ and $H$ to the whole of $\Omega$, which
we still denote by $f$, and $H$, respectively.
Letting $\varphi(r) = r^2$, using Green's
identity, and arguing as in section \ref{section_basic_energy},
\begin{align}
\begin{split}
\int_\Omega \left( \varphi \overline{\Delta} H \Delta f - f \Delta( \varphi \overline{\Delta} H )
\right)
& = \int_{\partial \Omega} \left( \overline{\Delta} H \partial_\nu f - f \partial_\nu (\varphi
\overline{\Delta} H ) \right)
= \int_{\partial \Omega} \left( \overline{\Delta} H \partial_\nu f -
f \overline{\Delta} \partial_\nu H \right),
\end{split}
\label{integration_parts_spatial_H_f}
\end{align}
where we used (\ref{induced_Laplacian_spherical}) so that, on $\partial \Omega$,
\begin{gather}
\partial_\nu (\varphi\overline{\Delta} H ) = \partial_r(r^2 \overline{\Delta} H )
= \partial_r \Delta_{S^2} H = \Delta_{S^2} \partial_r H =
\overline{\Delta} \partial_\nu H.
\nonumber
\end{gather}
Using also (\ref{Laplacian_spherical}),
\begin{align}
\begin{split}
\Delta (\varphi \overline{\Delta} H ) & =
\Delta_{S^2} \partial_r^2 H
+ \Delta_{S^2}\left( \frac{2}{r} \partial_r H \right)
+ \Delta_{S^2} \left (\frac{1}{r^2} \Delta_{S^2} H \right)
= \Delta_{S^2} \Delta H = 0.
\end{split}
\nonumber
\end{align}
And as $\Delta f = 0$, (\ref{integration_parts_spatial_H_f}) gives
\begin{gather}
\int_{\partial \Omega} \overline{\Delta} H \partial_\nu f =
\int_{\partial \Omega} f \overline{\Delta} \partial_\nu H.
\label{integration_parts_spatial_H_f_result}
\end{gather}
Using (\ref{integration_parts_time_H_f}) and (\ref{integration_parts_spatial_H_f_result}) into
(\ref{integration_parts_L2_T}) yields,
\begin{gather}
(H, \ddot{f} - \overline{\Delta} \partial_\nu f )_{L^2(T)}
= (\ddot{H} - \overline{\Delta} \partial_\nu H, f )_{L^2(T)}.
\nonumber
\end{gather}
Since this holds for all $ f \in X^{3,+}_T(\partial \Omega)$ we conclude, from
(\ref{inner_product_zero_1})
and the density
of $X^{3,+}_T(\partial \Omega)$ in $L^2(T)$, that
$(\ddot{H} - \overline{\Delta} \partial_\nu H, f )_{L^2(T)} = 0$ for all $f \in L^2(T)$,
and thus
\begin{gather}
\ddot{H} - \overline{\Delta} \partial_\nu H = 0.
\nonumber
\end{gather}
Recall that $H(T) = \dot{H}(T) = 0$ because $H \in X^{3,-}_T(\partial \Omega)$.
But solutions to the above equation, with these boundary conditions, are unique;
this follows using the same argument used to show that $\mathcal L^{-1}$
is injective, i.e., using energy estimates, except
with the roles of $0$ and $T$ reversed.
Thus $H = 0$. Since $X^{3,-}_T(\partial \Omega)$ is dense in $L^2(T)$, we conclude from
the continuity of the inner product that if $(v,w)_\mathcal H = 0$ for all $w \in \mathcal R$, then
$v = (0, v_0, v_1)$.
(\ref{inner_product_zero_1}) therefore reduces to
\begin{gather}
(v_0, f(0))_{\frac{3}{2}, \partial} + (v_1, \dot{f}(0))_{0,\partial} = 0
\text{ for all } f \in X^3_T(\partial \Omega).
\nonumber
\end{gather}
But if $f$ is an arbitrary element of $X^3_T(\partial \Omega)$, then $f(0)$ and
$\dot{f}(0)$ are arbitrary elements of $H^3(\partial \Omega)$ and $H^\frac{3}{2}(\partial \Omega)$,
respectively. Since these last two spaces are dense in, respectively,
$H^\frac{3}{2}(\partial \Omega)$ and $H^0(\partial \Omega)$, we conclude that
\begin{gather}
(v_0, f_0)_{\frac{3}{2}, \partial} + (v_1, f_1)_{0,\partial} = 0
\text{ for all } (f_0, f_1) \in H^\frac{3}{2}(\partial \Omega) \times H^0(\partial \Omega),
\nonumber
\end{gather}
and thus $v = 0$, as desired. \qed
\noindent \emph{Proof of proposition \ref{prop_extension}-(iii):}
We already know that $\overline{\mathcal L^{-1}}$ is defined on the
whole of $\mathcal H$, and any $u \in \mathcal H$ is the limit of a sequence in $\mathcal R$.
Let $ y = \overline{\mathcal L^{-1}}(u)$, and take a sequence $\{ u_\ell \} \subset \mathcal R$
converging in $\mathcal H$ to $u$. Then
\begin{gather}
\overline{\mathcal L^{-1}}(u_\ell ) = \mathcal L^{-1}(u_\ell) \equiv y_\ell.
\nonumber
\end{gather}
By construction, $\mathcal L^{-1}$ is the inverse of $\mathcal L$ defined on $X^3_T(\partial \Omega)$, thus
$y_\ell \in X^3_T(\partial \Omega)$. As we showed that $\mathcal L^{-1}$ is a continuous map
(see (\ref{continuity_L_inverse})), we have
\begin{align}
\begin{split}
\parallel y - y_\ell \parallel_{X^\frac{3}{2}_T(\partial \Omega)}
=
\parallel \overline{\mathcal L^{-1}}(u) - \mathcal L^{-1}(u_\ell) \parallel_{X^\frac{3}{2}_T(\partial \Omega)}
=
\parallel \overline{\mathcal L^{-1}}(u - u_\ell) \parallel_{X^\frac{3}{2}_T(\partial \Omega)}
\leq C \parallel u - u_\ell \parallel_\mathcal H,
\end{split}
\nonumber
\end{align}
implying that $y_\ell \rightarrow y$ in $X^\frac{3}{2}_T(\partial \Omega)$, and thus
$y \in \overline{X^3_T(\partial \Omega)\,}{}^{X^\frac{3}{2}_T(\partial \Omega)}$.
Injectivity also follows from the continuity of $\overline{\mathcal L^{-1}}$.
Say $\overline{\mathcal L^{-1}}(u) = 0$, and let $\{ u_\ell \}$ be as above.
Then
\begin{gather}
0 = \overline{\mathcal L^{-1}}(u) = \overline{\mathcal L^{-1}}(\lim u_\ell)
= \lim \overline{\mathcal L^{-1}}(u_\ell),
\nonumber
\end{gather}
which implies $u_\ell = 0$ for every $\ell$ since
$\overline{\mathcal L^{-1}} = \mathcal L^{-1}$ on $\mathcal R$ and $\mathcal L^{-1}$ is injective.
Thus $u = 0$. \qed \\
\noindent \emph{Proof of theorem \ref{main_theorem}:}
The existence and uniqueness of a weak solution follows at once from
proposition \ref{prop_extension} , upon noticing that if
$f \in X_T^\frac{3}{2}(\partial \Omega)$, then, by elliptic theory,
its harmonic extension
is in $X_T^2( \Omega)$. \qed
|
\section{Introduction}
Drell--Yan-like W- or Z-boson production is among the most important
standard candle processes at the LHC.
Apart from delivering important information on parton distributions
and allowing for the search for new gauge bosons in the high-mass range,
these processes allow for high-precision measurements in the resonance regions.
The weak mixing angle might be measured with LEP precision,
and the W-boson mass $\mathswitch {M_\PW}$ with an accuracy exceeding
$10\unskip\,\mathrm{MeV}$.
In the past two decades, great effort was made in the theory community
to deliver precise predictions matching the required accuracy
(for a list of references, see \citere{Dittmaier:2014qza}).
QCD corrections are known up to next-to-next-to-leading order,
electroweak (EW) corrections up to next-to-leading order (NLO).
Both on the QCD and on the EW side, there are further refinements such as
leading higher-order effects, resummations, matched parton showers.
In view of fixed-order calculations, the largest missing piece seems to be
the mixed QCD--EW corrections of ${\cal O}(\alpha_{\mathrm{s}}\alpha)$.
Knowing the contribution of this order will also answer the question how
to properly combine QCD and EW corrections in predictions.
In \citere{Balossini:2009sa} this issue is quantitatively discussed with special
emphasis on observables that are relevant for the $\mathswitch {M_\PW}$ determination,
revealing percent corrections of ${\cal O}(\alpha_{\mathrm{s}}\alpha)$
that should be calculated.
First steps towards this direction have been taken
by calculating two-loop contributions~\cite{Kotikov:2007vr},
the full ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ correction to the W/Z-decay widths~\cite{Czarnecki:1996ei},
and the full ${\cal O}(\alpha)$ EW corrections to W/Z+jet production
including the W/Z decays~\cite{Denner:2009gj}.
In this short article, we briefly report on our effort~\cite{Dittmaier:2014qza}
to calculate the ${\cal O}(\alpha_{\mathrm{s}}\alpha)$
corrections to Drell--Yan processes in the resonance region via the
so-called {\it pole approximation}, which is based on a systematic expansion
about the resonance pole.
Specifically, we sketch the salient features of the approach, discuss its success
at NLO, and describe results on the so-called non-factorizable contributions
at ${\cal O}(\alpha_{\mathrm{s}}\alpha)$, which comprise the most delicate
contribution to the PA.
Moreover, we present first (preliminary) results on the dominant factorizable
corrections, originating from the interplay of
initial-state QCD and final-state electroweak corrections.
\section{Pole approximation for NLO corrections}
The general idea~\cite{Stuart:1991xk}
of a pole approximation (PA) for any Feynman diagram with a single
resonance is the systematic isolation of all parts that are enhanced by
a resonance factor $1/(p^2-M_V^2+{\mathrm{i}}\mathswitch {M_\PV}\Gamma_V)$, where
$p$, $\mathswitch {M_\PV}$, and $\Gamma_V$ are the momentum, mass, and width
of the resonating particle~$V$, respectively.
In Drell--Yan production, $V$ stands for a W or a Z~boson.
For W~production different variants of PAs have been
suggested and discussed at NLO already in
\citeres{Wackeroth:1996hz,Dittmaier:2001ay}.
For the virtual corrections we follow the PA approach of \citere{Dittmaier:2001ay}.
Note, however, that we apply the PA to the real corrections as well, in contrast to
\citere{Dittmaier:2001ay} where they were based on full matrix elements.
Schematically each transition amplitude has the form
\begin{equation}
{\cal{M}}=\frac{W(p^2)}{p^2-M_V^2+\Sigma(p^2)}+N(p^2),
\end{equation}
with functions $W$ and $N$ describing resonant and non-resonant parts, respectively,
and $\Sigma$ denoting the self-energy of $V$.
The resonance of ${\cal{M}}$ is isolated in a gauge-invariant way as follows,
\begin{equation}
{\cal{M}} =
\frac{W(\mu_V^2)}{p^2-\mu_V^2}\,\frac{1}{1+\Sigma'(\mu_V^2)}
+\left[\frac{W(p^2)}{p^2-M_V^2+\Sigma(p^2)}
- \frac{W(\mu_V^2)}{p^2-\mu_V^2}\,\frac{1}{1+\Sigma'(\mu_V^2)} \right] + N(p^2),
\label{eq:PA}
\end{equation}
where $\mu_V^2=M_V^2-{\mathrm{i}} M_V\Gamma_V$ is the gauge-invariant location of the
propagator pole in the complex $p^2$ plane.
Equation~\refeq{eq:PA} can serve as a basis for the gauge-invariant introduction of
the finite decay width in the resonance propagator, thereby defining the so-called
{\it pole scheme}. In this scheme the term in square brackets is
perturbatively expanded in the coupling $\alpha$ including terms up to ${\cal O}(\alpha)$,
while the full $p^2$ dependence is kept.
An application of this scheme to Z-boson production is, e.g.,
described in \citere{Dittmaier:2009cr} in detail.
The PA for the amplitude
results from the r.h.s.\ of \refeq{eq:PA} upon neglecting the last, non-resonant term and
asymptotically expanding the term in square brackets
in $p^2$ about the point $p^2=\mu_V^2$,
where only the leading, resonant term of the expansion is kept.
The first term on the r.h.s.\ of \refeq{eq:PA} defines the so-called
{\it factorizable} corrections in which on-shell production and decay amplitudes
for $V$ are linked by the off-shell propagator; these contributions
are illustrated by diagrams (b) and (c) of \reffi{fig:vNLOgraphs}.
\begin{figure}
\epsfig{file=DY_B.eps,scale=0.75}
\hfill
\epsfig{file=DY_NLO_fac_p.eps,scale=0.75}
\hfill
\epsfig{file=DY_NLO_fac_d.eps,scale=0.75}
\hfill
\epsfig{file=DY_NLO_nf_gen.eps,scale=0.75}
\\[-.5em]
\hspace*{3.5em} (a) \hfill (b) \hfill (c) \hfill (d) \hspace*{4em}
\caption{Generic diagrams for the lowest-order amplitude (a),
for the EW virtual NLO factorizable corrections to production (b) and decay (c),
as well as for virtual non-factorizable corrections (d), where the empty blobs stand for all relevant
tree structures and the ones with ``$\alpha$'' inside
for one-loop corrections of ${\cal O}(\alpha)$.}
\label{fig:vNLOgraphs}
\end{figure}
The term on the r.h.s.\ of \refeq{eq:PA} in square brackets contains the so-called
{\it non-factorizable} corrections which receives resonant contributions from all diagrams
where the limit $p^2\to\mu_V^2$ in $W(p^2)$ or $\Sigma'(p^2)$ would lead to (infrared)
singularities. At NLO, this happens if a soft photon of energy $E_\gamma\lsim\Gamma_V$
is exchanged between the production process, the decay part, and the intermediate $V$ bosons;
a generic loop diagram is shown in \reffi{fig:vNLOgraphs}(d).
Although in principle the real-emission corrections can be based on full amplitudes
without further approximations,
the consistent application of the PA to both virtual and real corrections
is necessary to make a separate discussion of factorizable
and non-factorizable contributions possible.
\looseness -1
Conceptually the evaluation of the non-factorizable corrections is the
most delicate among all PA contributions. They possess rather interesting features.
When the invariant mass of the resonance is integrated over, their contribution
vanishes~\cite{Fadin:1993dz},
i.e.\ they only tend to distort the resonance without changing the normalization
of the cross section.
Since they only involve soft photons, they take the form of a global correction
factor to the lowest-order matrix element squared, with a non-trivial dependence
on the virtuality $(p^2-\mu_V^2)$ which gave the corrections their name.
The virtual correction factor can be explicitly calculated with quite
compact results, even for double resonances for which their calculation is described
in detail in the literature~\cite{Melnikov:1995fx}.
The real corrections are better evaluated numerically to keep some flexibility
in the treatment of photons in the event selection, using extended
eikonal currents~\cite{Melnikov:1995fx}
that take into account the resonance distortion by soft photons
of energy $E_\gamma\lsim\Gamma_V$.
Collinear singularities do not occur at all, since all relevant diagrams
are of interference type.
In spite of the simple general idea of the PA, its consistent implementation
in higher-order calculations involves subtle details.
For instance, care has to be taken that subamplitudes appearing before or after
the $V$ resonance are based on subamplitudes with on-shell $V$ bosons,
otherwise gauge invariance cannot be guaranteed.
Setting $p^2=\mathswitch {M_\PV}^2$, instead of the problematic complex value $p^2=\mu_V^2$,
in the ${\cal O}(\alpha)$ corrections is certainly allowed in ${\cal O}(\alpha)$ approximation.
However, the procedure is not unique, because the phase space is parametrized by
more than one variable. The {\it on-shell projection} $p^2\to\mathswitch {M_\PV}^2$ has to be defined
carefully. Different variants may lead to results that differ within the
intrinsic uncertainty of the PA, which is of ${\cal O}(\alpha/\pi\times\Gamma_V/\mathswitch {M_\PV})$
in the resonance region when applied to ${\cal O}(\alpha)$ corrections.
However, care has to be taken that virtual and real corrections still match properly
in the (soft and collinear) infrared limits in order to guarantee the cancellation
of the corresponding singularities.
Based on our results derived in \citere{Dittmaier:2014qza},
Figure~\ref{fig:PA-NLO} exemplarily shows the NLO QCD and EW corrections to the
transverse-mass and transverse-lepton-momentum distributions for $\mathswitchr {W^+}$ production
at the LHC and, in particular, illustrates the structure and quality of the PA
applied to the EW corrections.
\begin{figure}
\epsfig{file=Wp.mtll.abs.eps,scale=0.7} \hfill \epsfig{file=Wp.ptl.abs.eps,scale=0.7} \hfill
\\[.5em]
\hspace*{1.5em}\epsfig{file=Wp.mtll.rel.PA.eps,scale=0.7} \hfill \epsfig{file=Wp.ptl.rel.PA.eps,scale=0.7} \hfill
\caption{Distributions in the transverse mass (left) and transverse lepton momentum (right)
for $\mathswitchr {W^+}$ production at the LHC, with the upper plots showing the absolute distributions
and the lower plots the relative NLO
EW corrections in PA broken up into its factorizable and non-factorizable parts
(taken from \citere{Dittmaier:2014qza}).}
\label{fig:PA-NLO}
\end{figure}
The distributions show the well-known Jacobian peaks at $M_{{\mathrm{T}},\nul}\sim\mathswitch {M_\PW}$ and
$p_{{\mathrm{T}},l}\sim\mathswitch {M_\PW}/2$, respectively, which play a central role in the measurement of
the W-boson mass $\mathswitch {M_\PW}$ at hadron colliders.
The EW corrections significantly distort the distributions and shift the peak position.
Note also the extremely large QCD corrections above the peak in the $p_{{\mathrm{T}},l}$
distribution, which are induced by the recoil of the W~boson against the hard jet
of the real QCD correction.
The lower panels of \reffi{fig:PA-NLO} compare the full NLO EW corrections
(without photon-induced processes from $q\gamma$ collisions)
to the result of the PA, which is also broken up into factorizable corrections to
the initial/final state and non-factorizable contributions.
Near the Jacobian peaks, the PA turns out to be good within some $0.1\%$.
Interestingly, the impact of the non-factorizable corrections is suppressed to the
$0.1\%$ level and, thus, phenomenologically negligible. A similar conclusion
even holds for the factorizable initial-state corrections when one takes into account that
the percentage correction to the $p_{{\mathrm{T}},l}$ distribution actually should be normalized
to the full cross section including QCD corrections, which are overwhelming above
the Jacobian peak.
Thus, the relevant part of the NLO EW corrections near the peaks entirely results
from the factorizable final-state corrections, where the bulk originates from
collinear final-state radiation from the decay leptons.
More details and results on the PA at NLO
are discussed in \citere{Dittmaier:2014qza}, in particular for $\mathswitchr Z$~production, for which
the PA works similarly well.
\section{Pole approximation at \boldmath{${\cal O}(\alpha_{\mathrm{s}}\alpha)$}}
Though technically more complicated,
the concept of the PA, described at NLO in the previous section, can be carried over
to higher orders in a straightforward way.
The corresponding virtual, real, and mixed virtual--real
contributions involve various different interference diagrams.
Exemplarily we depict the generic graphs for the non-factorizable
${\cal O}(\alpha_{\mathrm{s}}\alpha)$ corrections in \reffi{fig:nfOaasgraphs}
and for the factorizable contributions combining initial-state QCD and final-state EW
corrections in \reffi{fig:IFfactOaasgraphs}.
\looseness-1
\begin{figure}
\centerline{\epsfig{file=DY_MIX_vv1.eps,scale=0.75}
\hspace*{3em}
\epsfig{file=DY_MIX_vv2.eps,scale=0.75}}
\centerline{\epsfig{file=DY_MIX_rv1.eps,scale=0.75}
\hspace*{3em}
\epsfig{file=DY_MIX_rv2.eps,scale=0.75}}
\centerline{\epsfig{file=DY_MIX_vr1.eps,scale=0.75}
\hspace*{3em}
\epsfig{file=DY_MIX_vr2.eps,scale=0.75}}
\centerline{\epsfig{file=DY_MIX_rr1.eps,scale=0.75}
\hspace*{3em}
\epsfig{file=DY_MIX_rr2.eps,scale=0.75}}
\vspace*{-.5em}
\caption{Generic diagrams for the non-factorizable
corrections of ${\cal O}(\alpha_{\mathrm{s}}\alpha)$.}
\label{fig:nfOaasgraphs}
\end{figure}
\begin{figure}
\centerline{\epsfig{file=DY_MIXif_vv1.eps,scale=0.75}
\hspace*{3em}
\epsfig{file=DY_MIXif_vv2.eps,scale=0.75}}
\centerline{\epsfig{file=DY_MIXif_rv1.eps,scale=0.75}
\hspace*{3em}
\epsfig{file=DY_MIXif_rv2.eps,scale=0.75}}
\centerline{\epsfig{file=DY_MIXif_vr1.eps,scale=0.75}}
\centerline{\epsfig{file=DY_MIXif_rr1.eps,scale=0.75}
\hspace*{3em}
\epsfig{file=DY_MIXif_rr2.eps,scale=0.75}}
\vspace*{-.5em}
\caption{Generic diagrams for the ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ factorizable
corrections of the ``initial--final'' type.}
\label{fig:IFfactOaasgraphs}
\end{figure}
As at NLO, the non-factorizable corrections originate from the virtual exchange or
real emission of soft photons with energies $E_\gamma\lsim\Gamma_V$, however, without
any restriction on the kinematics of virtual gluons or real jet radiation.
In \citere{Dittmaier:2014qza} we have discussed in detail the factorization properties of the virtual and real
photonic parts
of the non-factorizable ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ corrections, which result
from the soft nature of the effect.
Using gauge-invariance arguments borrowed from the classic YFS paper~\cite{Yennie:1961ad},
we show that this factorization of the photonic factors even hold to any order in the
strong coupling $\alpha_{\mathrm{s}}$.
We have verified this statement diagrammatically and, for the
purely virtual corrections, also with effective-field-theory
techniques~\cite{Beneke:2003xh}.
Both the virtual and real photonic corrections can be written as correction factors to
squared matrix elements containing gluon loops or external gluons, i.e.\ the necessary
building blocks are obtained from tree-level and one-loop calculations.
Our numerical study~\cite{Dittmaier:2014qza} shows that the non-factorizable corrections
of ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ below the $0.1\%$ level and, thus, phenomenologically negligible,
both for W-boson and Z-boson production.
Of course, one could have speculated on this suppression, since the impact
of non-factorizable photonic corrections is already at the level of some $0.1\%$
at NLO. However, the ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ corrections mix EW and QCD
effects, so that small photonic corrections might have been enhanced
by the strong jet recoil effect observed in the $p_{{\mathrm{T}},l}$ distribution.
This enhancement is seen in the virtual and real corrections separately, but not in their sum.
Furthermore, the existence of gluon-induced ($q\mathswitchr g$) channels implies
a new feature in the non-factorizable corrections. In the $q\bar q$ channels,
and thus in the full NLO part of the non-factorizable corrections,
the soft-photon exchange proceeds between initial- and final-state particles, whereas
in the $q\mathswitchr g$ channels at ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ the photon is also
exchanged between final-state particles. The known suppression mechanisms in
non-factorizable corrections work somewhat differently in those cases~\cite{Melnikov:1995fx}.
Figures~\ref{fig:Oaas-IFfact} and \ref{fig:Oaas-IFfact-2}
show first preliminary results on the
``initial--final'' factorizable ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ corrections
$\delta^{\mathrm{ini-fin}}_{\alpha_{\mathrm{s}}\alpha}$
induced by initial-state QCD and final-state EW contributions.
\begin{figure}
\hspace*{1.5em}\epsfig{file=Wp.mtll.Oaas-fact.rel.eps,scale=0.7} \hfill
\epsfig{file=Wp.ptl.Oaas-fact.rel.eps,scale=0.7}
\\[-2.2em]
\caption{Relative factorizable corrections (in red) of ${\cal O}(\alpha_{\mathrm{s}}\alpha)$
induced by initial-state QCD and final-state EW contributions
to the distributions in $M_{\mathrm{T},\nu l}$ (left) and $p_{\mathrm{T,l}}$ (right)
for $\mathswitchr {W^+}$ production at the LHC. The naive products of the NLO correction
factors $\delta_{\alpha_{\mathrm{s}}}$ and $\delta_{\alpha}$
are shown for comparison (see text).}
\label{fig:Oaas-IFfact}
\vspace*{1.5em}
\hspace*{1.5em}\epsfig{file=Wp.mtll.Oaas-fact.rel2.eps,scale=0.7} \hfill
\epsfig{file=Wp.ptl.Oaas-fact.rel2.eps,scale=0.7}
\\[-2.2em]
\caption{As in Fig.~5, but with the naive product of QCD and EW
corrections based on $\delta^{\prime}_{\alpha_{\mathrm{s}}}$ instead
\looseness -1
of~$\delta_{\alpha_{\mathrm{s}}}$.
}
\label{fig:Oaas-IFfact-2}
\end{figure}
From the results of the PA for the NLO EW corrections, one has to expect that those
contributions furnish the by far dominant part at ${\cal O}(\alpha_{\mathrm{s}}\alpha)$,
while the two other types of factorizable contributions of ``initial--initial'' and
``final--final'' type are much smaller.
In detail, the figures compare the factorizable initial--final
corrections (red curves) to different versions of
naive products of the NLO QCD and NLO EW correction factors.
To define the naive products, we write the NLO QCD cross section
$\sigma^{\mathrm{NLO}_{\mathrm{s}}}$ as
\begin{equation}
\sigma^{\mathrm{NLO}_{\mathrm{s}}}
\equiv \sigma^{\mathrm{LO}}(1+\delta_{\alpha_{\mathrm{s}}})
= \sigma^0+\sigma^{\mathrm{LO}}\left(\frac{\sigma^{\mathrm{LO}}-\sigma^0}{\sigma^{\mathrm{LO}}}+\delta_{\alpha_{\mathrm{s}}}\right)
\equiv\sigma^0+\sigma^{\mathrm{LO}}\delta'_{\alpha_{\mathrm{s}}},
\end{equation}
where $\sigma^{\mathrm{LO}}$ and $\sigma^0$ denote the LO cross section
evaluated with LO and NLO PDFs, respectively. The standard QCD $K$~factor is, thus,
given by $K^{\mathrm{NLO}_{\mathrm{s}}}=1+\delta_{\alpha_{\mathrm{s}}}$, and
$\delta'_{\alpha_{\mathrm{s}}}$ differs from $\delta_{\alpha_{\mathrm{s}}}$ by the LO prediction induced by the
transition from LO to NLO PDFs.
The EW correction factor
is used in two different versions, first based on the full NLO
correction ($\delta_\alpha$, black curves), and second based on the dominant EW final-state correction
of the PA ($\delta^{\mathrm{fin}}_\alpha$, blue curves). The
relative EW correction is always derived from the ratio of the NLO EW contribution
$\Delta\sigma^{\mathrm{NLO}_{\mathrm{ew}}}$ and the LO contribution $\sigma^0$ to the NLO
cross section, $\delta_\alpha=\Delta\sigma^{\mathrm{NLO}_{\mathrm{ew}}}/\sigma^0$,
so that the EW correction factors are practically independent of the PDF set.
An ansatz for the cross section $\sigma^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}_{\mathrm{naive\,fact}}$
based on the naive factorization of QCD and EW corrections
then reads
\begin{equation}
\sigma_{\mathrm{naive\,fact}}^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}
= \sigma^{\mathrm{NLO}_{\mathrm{s}}}(1+\delta_\alpha)
= \sigma^{\mathrm{LO}}(1+\delta_{\alpha_{\mathrm{s}}})(1+\delta_\alpha).
\end{equation}
This ansatz should approximate the
full NLO QCD+EW cross section improved by our calculated ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ correction
$\Delta\sigma^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}_{\mathrm{ini-fin}}$,
\begin{equation}
\sigma^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}
= \sigma^{\mathrm{NLO}_{\mathrm{s}}}+ \Delta\sigma^{\mathrm{NLO}_{\mathrm{ew}}}
+ \Delta\sigma^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}_{\mathrm{ini-fin}},
\end{equation}
which is consistently evaluated with NLO PDFs.
Figures~\ref{fig:Oaas-IFfact} and \ref{fig:Oaas-IFfact-2} compare the
${\cal O}(\alpha_{\mathrm{s}}\alpha)$ correction
$\delta^{\mathrm{ini-fin}}_{\alpha_{\mathrm{s}}\alpha}
=\Delta\sigma^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}_{\mathrm{ini-fin}}/\sigma^{\mathrm{LO}}$
with the two versions $\delta_{\alpha_{\mathrm{s}}}\delta_\alpha$
and $\delta^{\prime}_{\alpha_{\mathrm{s}}}\delta_\alpha$, respectively.
The corrections to the W-boson transverse-mass distribution are approximated
by the naive factorization $\delta^{\prime}_{\alpha_{\mathrm{s}}}\delta_\alpha$ quite well.
The preference of the factorization variant $\delta^{\prime}_{\alpha_{\mathrm{s}}}\delta_\alpha$
over $\delta_{\alpha_{\mathrm{s}}}\delta_\alpha$ can be interpreted upon inspecting the
difference between $\sigma^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}$ and
the factorized ansatz $\sigma_{\mathrm{naive\,fact}}^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}$,
\begin{equation}
\frac{\sigma^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}
-\sigma_{\mathrm{naive\,fact}}^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}}{\sigma^{\mathrm{LO}}}
=
\delta^{\mathrm{ini-fin}}_{\alpha_{\mathrm{s}}\alpha}
-\delta'_{\alpha_{\mathrm{s}}}\delta_\alpha,
\end{equation}
i.e.\ the validity of the factorization approximation
$\sigma_{\mathrm{naive\,fact}}^{\mathrm{NNLO}_{\mathrm{s\otimes ew}}}$
is signalled by a small difference between
$\delta^{\mathrm{ini-fin}}_{\alpha_{\mathrm{s}}\alpha}$ and
$\delta'_{\alpha_{\mathrm{s}}}\delta_\alpha$.
Naive factorization works for the $M_{\mathrm{T},\nu l}$ distribution, which
is rather insensitive to any recoil effect of the W~boson
with respect to additional jet emission.
The factorization ansatz fails, however,
for the distribution in the lepton transverse momentum,
which is sensitive
to the interplay between QCD and photonic real-emission effects.
It remains to be seen whether the
${\cal O}(\alpha_{\mathrm{s}}\alpha)$ corrections to the $p_{\mathrm{T,l}}$ distribution
can be reproduced by attaching a photonic shower, which is based on the asymptotics of collinear
photon emission, to a QCD-based prediction.
\looseness-1
The results shown in \reffis{fig:Oaas-IFfact} and \ref{fig:Oaas-IFfact-2}
are obtained by photon recombination, which combines collinear photons and leptons to
a quasi-particle as described in \citeres{Dittmaier:2001ay,Dittmaier:2009cr},
while the NLO EW results shown in \reffi{fig:PA-NLO} involve ``bare muons''
without any photon recombination.
Photon recombination restores the level of inclusiveness required by the KLN theorem
to imply a cancellation of the collinear singularity, which enhances the effect of final-state
radiation. Photon recombination typically reduces the size of photonic corrections by a factor of two.
\looseness-1
\section{Conclusions}
Here we have briefly summarized the main results of \citere{Dittmaier:2014qza}, where
we have shown how the ${\cal O}(\alpha_{\mathrm{s}}\alpha)$
corrections to Drell--Yan processes can be approximated by the leading term in an
expansion about the resonance pole. The quality of such an approximation achieved at
NLO strongly supports the expectation that this approach is sufficient for
observables that are dominated by the resonance, which include, e.g., the ones
relevant for precision determinations of the W-boson mass.
The pole approximation classifies corrections into factorizable and
non-factorizable contributions. Our results show that the latter, which
link production and decay subprocesses via soft-photon exchange, are phenomenologically
negligible. The phenomenologically relevant corrections, thus, are of
factorizable nature and can be attributed to corrections to the initial or final states.
Again, the pattern of contributions to the pole approximation at NLO gives
a clear picture on the expected hierarchy in higher orders.
At ${\cal O}(\alpha_{\mathrm{s}}\alpha)$, the bulk of corrections will be
contained in the combination of QCD corrections to the production and EW
corrections to the decay processes, with a particular enhancement induced by
final-state radiation off the charged leptons.
Here we have presented first preliminary results on those contributions,
revealing that the naive factorization of NLO QCD and NLO EW corrections approximates
the ${\cal O}(\alpha_{\mathrm{s}}\alpha)$ corrections to the W-boson transverse-mass distribution,
but not the lepton transverse-momentum distribution, which is particularly sensitive
to recoil effects of the W~boson.
The completion of our calculation of ${\cal O}(\alpha_{\mathrm{s}}\alpha)$
corrections and the discussion of their phenomenological implications are in progress.
|
\section{Proof of Theorem \ref{loc-reg-th-lcf} and Corollary \ref{reg-th-lcf}}
\setcounter{equation}{0}
\setcounter{theorem}{0}
In this section, we will prove Theorem
\ref{loc-reg-th-lcf} and Corollary \ref{reg-th-lcf} for nematic liquid crystal flows (\ref{lce}). The crucial step
is to establish an $\epsilon_0$-regularity criterion.
\blm\label{unqlc-lemma3.1}{\it There exists $\epsilon_0>0$ such that
if $(u,\nabla d)\in L^p_tL^q_x(P_1(0,1))$, for some $p\geq 2$ and
$q\geq n$ satisfying (\ref{serrin-condition}), is a weak
solution to (\ref{lce}) that satisfies
\beq\label{unqlc3.3}
\|u\|_{L^{p}_t L^{q}_x(P_1(0,1))}+\|\nabla
d\|_{L^{p}_t L^{q}_x(P_1(0,1)))}\leq\epsilon_0,
\eeq
then $(u,d)\in
C^{\infty}(P_{\frac{1}{16}}(0,1))$, and
\beq\label{unqlc3.4}
\|u\|_{L^{\infty}(P_{\frac{1}{16}}(0,1))}+\|\nabla
d\|_{L^{\infty}(P_{\frac{1}{16}}(0,1))}\leq C\epsilon_0.
\eeq
}\elm
Before proving this lemma, we need the following inequality, due to Serrin \cite{serrin}.
\blm\label{unq-serrin-lemma}{\it For any open set $U\subset\mathbb
R^n$ and any open interval $I\subset\mathbb R$, let $f$, $g$, $h\in
L^2_tH^1_x(U\times I)$ and $f\in L^p_tL^q_x(U\times I)$ with $3\le
n\le q\le +\infty$ and $2\le p\le +\infty$ satisfying
(\ref{serrin-condition}). Then
\beq\label{unq2.5}
\int_{U\times I}|f||g||\nabla h| \leq C\|\nabla h\|_{L^2(U\times I)}
\|g\|_{L^2_tH^1_x(U\times I)}^{\frac{n}{q}}
\left\{\int_I\|f\|_{L^q(\mathbb R^n)}^{p}\|g\|_{L^2(\mathbb R^n)}^2\,dt\right\}^{\frac{1}{p}},
\eeq where $C>0$ depends only on $n$. }\elm
\noindent{\bf Proof of Lemma \ref{unqlc-lemma3.1}}.
For any $(x,t)\in
P_{\frac{1}{2}}(0,1)$ and $0<r<\frac{1}{2}$, we have, by
(\ref{unqlc3.3}),
\beq\label{unqlc3.5}
\|u\|_{L^{p}_tL^q_x(P_r(x,t))}+\|\nabla
d\|_{L^{p}_tL^q_x(P_r(x,t))}\leq\epsilon_0.
\eeq
We will divide the proof into two claims.
\noindent{\bf Claim 1}. $\nabla d\in L^{\gamma}(P_{\frac{1}{2}}(0,1))$ for any $1<\gamma<\infty$, and
\begin{equation}
\|\nabla d\|_{L^{\gamma}(P_{\frac{1}{4}}(0,1))}\leq C(\gamma)\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}.
\end{equation}
To show it, let $d_1: P_r(x,t)\rightarrow \R^3$ solve
\beq\label{unqlc3.6}
\left\{
\begin{split}
\partial_t d_1-\Delta d_1&=0, \quad\mbox{in } P_r(x,t)\\
d_1&=d, \quad\mbox{on }\partial_pP_r(x,t).
\end{split}\right.
\eeq Set $d_2=d-d_1$. Multiplying (\ref{lce})$_3$ and
(\ref{unqlc3.6}) by $d_2$, subtracting the resulting equations and
integrating over $P_r(x,t)$, we obtain
\beq\label{unqlc3.7}
\begin{split}
&\sup\limits_{t-r^2\leq \tau\leq t}\int_{B_r(x)}|d_2|^2(\cdot,\tau)+2\int_{P_r(x,t)}|\nabla d_2|^2\\
\leq& C\int_{P_r(x,t)}(|u||d_2||\nabla d|+|\nabla d||d_2||\nabla d|)=J_1+J_2.
\end{split}
\eeq By (\ref{unq2.5}), the Poincar$\acute{\mbox{e}}$ inequality and
the Young inequality, we have
\begin{eqnarray*}
|J_1|
&\lesssim&\begin{cases}
\|\nabla d\|_{L^2(P_r(x,t))} \|\nabla d_2\|_{L^2(P_r(x,t))}^{\frac{n}{q}}
\left\{\int_{t-r^2}^t\|u\|_{L^{q}(B_{r}(x))}^{p}\|d_2\|_{L^2(B_r(x))}^2\,d\tau\right\}^{\frac{1}{p}},\ & p<+\infty\\
\|\nabla d\|_{L^2(P_r(x,t))}\|\nabla d_2\|_{L^2(P_r(x,t))}
\|u\|_{L^{\infty}_tL^n_x(P_r(x,t))},\ & p=+\infty,
\end{cases}\\
&\leq& \begin{cases}\frac{1}{2}\|\nabla
d_2\|_{L^2(P_r(x,t))}^{2}+C\epsilon_0\|\nabla
d\|_{L^2(P_r(x,t))}^2
+C\epsilon_0^{\frac{p}2}\|d_2\|_{L^\infty_tL^2_x(P_r(x,t))}^2,\ & p<+\infty\\
\frac{1}{2}\|\nabla
d_2\|_{L^2(P_r(x,t))}^{2}+C\epsilon_0\|\nabla
d\|_{L^2(P_r(x,t))}^2, \ & p=+\infty.
\end{cases}
\end{eqnarray*}
Similarly, for $J_2$, we have
$$
|J_2| \leq \begin{cases}\frac{1}{2}\|\nabla
d_2\|_{L^2(P_r(x,t))}^{2}+C\epsilon_0\|\nabla
d\|_{L^2(P_r(x,t))}^2
+C\epsilon_0^{\frac{p}2}\|d_2\|_{L^\infty_tL^2_x(P_r(x,t))}^2,\ & p<+\infty\\
\frac{1}{2}\|\nabla
d_2\|_{L^2(P_r(x,t))}^{2}+C\epsilon_0\|\nabla
d\|_{L^2(P_r(x,t))}^2, \ & p=+\infty.
\end{cases}
$$
Putting these estimates into (\ref{unqlc3.7}), applying (\ref{unqlc3.5}), and choosing sufficiently small $\epsilon_0$,
we have
\beq\label{unqlc3.12}
\int_{P_r(x,t)}|\nabla d_2|^2
\leq C \epsilon_0\|\nabla d\|_{L^2(P_r(x,t))}^2.
\eeq
This, combined with the standard estimate on $d_1$, implies that for any $\theta\in (0,1)$,
\beq\label{unqlc3.14}
(\theta r)^{-n}\int_{P_{\theta r}(x,t)}|\nabla d|^2\le
C\Big(\theta^2+\theta^{-n}\epsilon_0\Big) r^{-n}\int_{P_r(x,t)}|\nabla d|^2.
\eeq
By iterations, we obtain for any $(x,t)\in P_{\frac{1}{2}}(0,1)$,
$0<r\leq\frac{1}{2}$ and $0<\alpha<1$,
\beq\label{unqlc3.15}
\begin{split}
r^{-n}\int_{P_{r}(x,t)}|\nabla d|^2
\leq Cr^{2\alpha}\int_{P_1(0,1)}|\nabla d|^2.
\end{split}
\eeq
Hence $\nabla d\in\mathcal{M}^{2,2-2\alpha}(P_\frac12(0,1))$ and
\begin{equation}\label{Dd_morrey_estimate}
\|\nabla d\|_{\mathcal M^{2,2-2\alpha}(P_\frac12(0,1))}\le
C\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}.
\end{equation}
Now Claim 1 follows
by the same estimate of Riesz potentials between parabolic Morrey spaces
as in \cite{huang-wang} (Theorem 1.5) and \cite{lin-wang} (Lemma 2.1).
\medskip
\noindent{\bf Claim 2}. $u\in L^{\gamma}(P_{\frac{1}{4}}(0,1))$ for
any $1<\gamma<\infty$, and
\begin{equation} \label{integral_estimate_u}
\|u\|_{L^\gamma(P_\frac14(0,1))}\le C(\gamma)\|u\|_{L^p_tL^q_x(P_1(0,1))}.
\end{equation}
Let $\mathbb{E}^\gamma$ be the closure in
$L^{\gamma}(\R^n,\R^n)$ of all divergence-free vector fields with
compact supports. Let $\mathbb{P}:L^{2}(\R^n,\R^n)\rightarrow
\mathbb{E}^2$ be the Leray projection operator. It is well-known
that $\mathbb{P}$ can be extended to a bounded linear operator from
$L^\gamma(\R^n,\R^n)$ to $\mathbb{E}^{\gamma}$ for all
$1<\gamma<+\infty$. Let $\mathbb{A}=\mathbb{P}\Delta$ denote the Stokes
operator.
For any $(x,t)\in P_{\frac{1}{4}}(0,1)$ and
$0<r\leq \frac{1}{4}$, let $\eta\in C_0^\infty(P_{2r}(x,t))$ be such that
$0\le\eta\le 1$, $\eta\equiv 1$ on $P_r(x,t)$, $|\nabla\eta|\le 4r^{-1}$,
and $|\partial_t\eta|\le 16r^{-2}$. Let $(v,P^1):\mathbb R^n\times (0,1)\to\mathbb R^n\times \mathbb R$ solve
\begin{equation}\label{stokes}
\begin{cases}
\partial_t v-\Delta v+\nabla P^1=-\nabla\cdot\Big(\eta^2 (u\otimes u+\nabla d\otimes\nabla d-\frac12{|\nabla d|^2}\mathbb{I}_n)\Big) &\ {\rm{in}}\ \mathbb R^n\times (0,1)\\
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \nabla\cdot v= 0 &\ {\rm{in}}\ \mathbb R^n\times (0,1)\\
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ v=0 &\ {\rm{on}}\ \mathbb R^n\times \{0\}.
\end{cases}
\end{equation}
Define $w: P_r(x,t)\rightarrow \R^n$ by $w=u-v$. Then $w$ solves
the Stokes equation in $P_r(x,t)$:
\beq\label{unqlc3.17} \left\{
\begin{split}
\partial_t w-\Delta w+\nabla Q^1&=0 \quad\mbox{in } P_r(x,t)\\
\nabla\cdot w&=0\quad\mbox{in } P_r(x,t).
\end{split}\right.
\eeq
By the standard theory of linear Stokes' equations, we have that
$w\in C^\infty(P_r(x,t))$ and,
for any $\theta\in (0, 1)$,
\begin{equation}\label{stokes_est}
\|w\|_{L^p_tL^q_x(P_{\theta r}(x,t))}
\le C\theta\ \|w\|_{L^p_tL^q_x(P_r(x,t))}.
\end{equation}
To estimate $v$, we apply $\mathbb P$ to both sides of the equation (\ref{stokes})$_1$ to obtain
$$\partial_t v-\mathbb A v=-\mathbb P\nabla\cdot\Big(\eta^2 (u\otimes u+\nabla d\otimes\nabla d-\frac12{|\nabla d|^2}\mathbb{I}_n)\Big)
\ {\rm{in}}\ \mathbb R^n\times (0,1); \
v=0 \ {\rm{on}}\ \mathbb R^n\times \{0\}.$$
By the Duhamel formula, we have
\begin{equation}\label{duhamel}
v(t)=-\int_0^te^{-(t-\tau)\mathbb{A}}\mathbb{P}\nabla\cdot\Big(\eta^2(u\otimes u+\nabla d\otimes\nabla d-\frac12{|\nabla d|^2}\mathbb{I}_n)\Big)\,d\tau,
\ 0<t\le 1.
\end{equation}
Now we can apply Fabes-Jones-Riviere \cite{FJR} Theorem 3.1 (see also Kato \cite{kato} page 474, ($2.3'$)) to conclude
that $v\in L^p_tL^q_x(\mathbb R^n\times [0,1])$ and
\begin{eqnarray}\label{v-estimate}
\|v\|_{L^p_tL^q_x(\mathbb R^n\times [0,1])}
&\le& C(\|\eta u\|_{L^p_tL^q_x(\mathbb R^n\times [0,1])}^2+\|\eta \nabla d\|_{L^p_tL^q_x(\mathbb R^n\times [0,1])}^2)\nonumber\\
&\le& C\epsilon_0 (\|u\|_{L^p_tL^q_x(P_{2r}(x,t)))}+\|\nabla d\|_{L^p_tL^q_x(P_{2r}(x,t))}).
\end{eqnarray}
Putting (\ref{stokes_est}) and (\ref{v-estimate}) together, we have that for any $\theta\in (0, 1)$,
\begin{equation}\label{decay_est1}
\|u\|_{L^p_tL^q_x(P_{\theta r}(x,t))}
\le C(\theta+\epsilon_0)\|u\|_{L^p_tL^q_x(P_{2r}(x,t))}
+C\epsilon_0\|\nabla d\|_{L^p_tL^q_x(P_{2r}(x,t))}.
\end{equation}
By Claim 1, we have that
for any $\alpha\in (0,1)$, there exists $\epsilon_0>0$ depending on $\alpha$ such that
\begin{equation}\label{Dd_estimate}
\|\nabla d\|_{L^p_tL^q_x(P_{2r}(x,t))}\le Cr^{\alpha}\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}.
\end{equation}
Substituting (\ref{Dd_estimate}) into (\ref{decay_est1}) yields
\begin{equation}\label{decay_est2}
\|u\|_{L^p_tL^q_x(P_{\theta r}(x,t))}
\le C(\theta+\epsilon_0)\|u\|_{L^p_tL^q_x(P_{2r}(x,t))}+Cr^\alpha\|\nabla d\|_{L^p_tL^q_x(P_{1}(0,1))}.
\end{equation}
It is standard that by choosing $\theta=\theta_0(\alpha)>0$ and
iterating (\ref{decay_est2}) finitely many times, we conclude that
for any $(x,t)\in P_{\frac{1}{4}}$,
$0<r\leq\frac{1}{4}$ and $0<\alpha<1$,
\begin{equation}\label{decay_est3}
\|u\|_{L^p_tL^q_x(P_{r}(x,t))}
\le C\Big(\|u\|_{L^p_tL^q_x(P_1(0,1))}+\|\nabla d\|_{L^p_tL^q_x(P_{1}(0,1))}\Big)r^\alpha.
\end{equation}
By H\"older's inequality, (\ref{decay_est3}) implies that
$u\in \mathcal M^{2,2-2\alpha}(P_\frac38(0,1))$, and
\begin{equation}\label{u_morrey_estimate}
\|u\|_{\mathcal M^{2,2-2\alpha}(P_\frac38(0,1))}
\le C\Big[\|u\|_{L^p_tL^q_x(P_1(0,1))}+\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}\Big].
\end{equation}
The higher integrability estimate of $u$ on $P_\frac14(0,1)$ can be done by the parabolic Riesz potential estimate
in parabolic Morrey spaces. Here we will sketch it.
Let $\phi\in C_{0}^{\infty}(P_{\frac{3}{8}}(0,1))$ such that $0\leq\phi\leq 1$, $\phi\equiv1$ on
$P_{\frac{5}{16}}(0,1)$, and
$$|\partial_t\phi|+|\nabla \phi|+|\nabla^2\phi|\leq C.$$
Define $\widetilde{u}:\mathbb R^n\times [0,1]\to\mathbb R^n$ by
\begin{equation}\label{duhamel1}
\widetilde{u}(t)=-\int_0^te^{-(t-\tau)\mathbb{A}}\mathbb{P}\nabla\cdot\Big(\phi^2(u\otimes u+\nabla d\otimes\nabla d-\frac12{|\nabla d|^2}\mathbb{I}_n)\Big)\,d\tau,
\ 0<t\le 1.
\end{equation}
Then, as in the proof of Theorem 3.1 (i) of \cite{FJR}, we have that for any $(x,t)\in\mathbb R^n\times (0,1]$,
\begin{equation}\label{duhamel2}
|\widetilde{u}(x,t)|
\le C\int_0^t\int_{\mathbb R^n}\frac{1}{\delta^{n+1}((x,t), (y,s))}
(|\phi u|^2+|\phi\nabla d|^2)(y,s)\,dyds.
\end{equation}
Recall the parabolic Riesz potential of
order $1$, $I_1(\cdot)$, is defined by
$$I_{1}(f)(z):=\int_{\R^{n+1}}\frac{|f(w)|}{\delta^{n+1}(z,w)}\, dw, \ f\in L^1(\mathbb R^{n+1}).$$
Then we have
\begin{equation}\label{para_morrey}
|\widetilde{u}(x,t)|
\le CI_1({F})(x,t), \ (x,t)\in\mathbb R^n\times (0,1],
\end{equation}
where
$${F}=\phi^2(|u|^2+|\nabla d|^2).$$
By H\"older's inequality, (\ref{Dd_morrey_estimate}), and (\ref{u_morrey_estimate}), we have
that ${F}\in \mathcal{M}^{1,2-2\alpha}(\R^{n+1})$ and
\begin{equation}\label{f_estimate}
\|{F}\|_{\mathcal M^{1,2-2\alpha}(\mathbb R^{n+1})}
\le C\Big(\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}^2+\|u\|_{L^p_tL^q_x(P_1(0,1))}^2\Big).
\end{equation}
Hence, by \cite{huang-wang} Theorem 3.1 (ii), we conclude that
$\widetilde{u}\in \mathcal M_*^{\frac{2-2\alpha}{1-2\alpha},2-2\alpha}(\R^{n}\times [0,1])$,
and
\begin{eqnarray}\label{morrey_estimate_u}
\|\widetilde{u}\|_{\mathcal M_*^{\frac{2-2\alpha}{1-2\alpha},2-2\alpha}(\R^{n}\times [0,1])}
&\le& C\|{F}\|_{\mathcal{M}^{1,2-2\alpha}(\R^{n+1})}\nonumber\\
&\le& C\left(\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}^2+\|u\|_{L^p_tL^q_x(P_1(0,1))}^2\right).
\end{eqnarray}
As $\lim\limits_{\alpha\uparrow \frac{1}{2}}\frac{2-2\alpha}{1-2\alpha}=+\infty$,
we have that $\widetilde{u}\in L^{\gamma}(P_{\frac{5}{16}}(0,1))$ for any
$1<\gamma<+\infty$, and
\begin{equation}\label{integral_estimate}
\|\widetilde u\|_{L^\gamma(P_\frac{5}{16})}\le C(\gamma)\Big(\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}^2+\|u\|_{L^p_tL^q_x(P_1(0,1))}^2\Big).
\end{equation}
Set $\widetilde w=u-\widetilde u$ on $P_\frac{5}{16}(0.1)$. Then it follows from (\ref{lce}) and (\ref{duhamel1})
that
$$\partial_t \widetilde w-\Delta\widetilde w+\nabla \widetilde{Q}=0;
\ \nabla\cdot\widetilde w=0\ \ \ {\rm{in}} \ \ P_{\frac{5}{16}}(0,1).$$
By the standard theory of linear Stokes' equations, we have that
$\widetilde w\in L^\infty(P_\frac14(0,1))$, and
\begin{eqnarray}\label{integral_estimate1}
\|\widetilde w\|_{L^\infty(P_\frac14(0,1))}&\le&
C\|\widetilde w\|_{L^1(P_\frac5{16}(0,1))}
\le C\Big(\|u\|_{L^1(P_\frac5{16}(0,1))}+\|\widetilde u\|_{L^1(P_\frac5{16}(0,1))}\Big)\nonumber\\
&\le& C\Big(\|\nabla d\|_{L^p_tL^q_x(P_1(0,1))}+\|u\|_{L^p_tL^q_x(P_1(0,1))}\Big).
\end{eqnarray}
It is clear that (\ref{integral_estimate_u}) follows from (\ref{integral_estimate}) and (\ref{integral_estimate1}).
This completes the proof of Claim 2.
Finally, it is not hard to see that by the $W^{2,1}_\gamma$-theory for the heat equation and the linear Stokes equation,
and the Sobolev embedding theorem, we have that $(u,\nabla d)\in L^\infty(P_\frac18(0,1))$. Then the Schauder's theory
and the bootstrap argument can imply that $(u,d)\in C^\infty(P_\frac1{16}(0,1))$.
Furthermore, the estimate (\ref{unqlc3.4}) holds.
This completes the proof. \endpf
\vspace{2mm}
\noindent{\bf Proof of Corollary \ref{reg-th-lcf}}: It is easy to see that when $p> 2$, $q>n$, for any $(x,t)\in\mathbb R^n\times(0,T]$, we can find $R_0>0$ such that
\beq\label{u-lcf-3}
\|u\|_{L^p_t L^q_x(P_{R_0}(x,t))}
+\|\nabla d\|_{L^p_t L^q_x(P_{R_0}(x,t))}\leq \epsilon_0,
\eeq
where $\epsilon_0$ is given in Lemma \ref{unqlc-lemma3.1}. By Theorem \ref{loc-reg-th-lcf}, we conclude that
$(u,d)\in C^{\infty}(P_{\frac{R_0}{16}}(x,t))$. This completes the proof of Theorem \ref{reg-th-lcf}
\endpf
\section{Proof of Theorem \ref{unique3}}
\setcounter{equation}{0}
\setcounter{theorem}{0}
\medskip
In this section, we will prove Theorem \ref{unique3}.
To do this, we need the
following estimate.
\blm\label{unqlc-lemma3.2}{For $T>0$, suppose that $(u,d)$ is a
weak solution to (\ref{lce}) in $\mathbb R^n\times (0,T]$, which
satisfies the assumption of Theorem \ref{unique3}. Then $(u,d)\in
C^{\infty}(\R^n\times(0,T],\mathbb R^n\times S^2)$, and there exists $t_0>0$ such that
for $0<t\le t_0$, it holds
\beq\label{unqlc3.32} \sup\limits_{0<\tau\leq
t}\sqrt{\tau}\Big(\|u(\tau)\|_{L^{\infty}(\R^n)}+\|\nabla
d(\tau)\|_{L^{\infty}(\R^n)}\Big)\leq C\Big(\|u\|_{L^p_t L^q_x(\R^n\times [0,t])}
+\|\nabla d\|_{L^p_t L^q_x(\R^n\times [0,t])}\Big).
\eeq
In particular, we have
\beq\label{unqlc3.36}
\lim\limits_{t\downarrow
0^+}\sqrt{t}\Big(\|u\|_{L^{\infty}(\R^n)}+\|\nabla
d\|_{L^{\infty}(\R^n)}\Big)=0. \eeq}
\elm
\pf Let $\epsilon_0$ be given by Lemma \ref{unqlc-lemma3.1}.
Since $p>2$ and $q>n$ satisfy (\ref{serrin-condition}),
for any $0<\epsilon\leq\epsilon_0$ we can find $t_0>0$ such that for any $0<\tau\leq \sqrt{t_0}$
\beq\label{u-lcf-3}
\|u\|_{L^p_t L^q_x(\mathbb R^n\times[0,\tau^2])}
+\|\nabla d\|_{L^p_t L^q_x(\mathbb R^n\times[0,\tau^2])}\leq \epsilon.
\eeq
For any $x_0\in \mathbb R^n$, define
\beq\notag
\begin{split}
&\bar{u}(y,s)=\tau u(x_0+y\tau ,s\tau ^2)\\
&\bar{P}(y,s)=\tau^2P(x_0+y\tau ,s\tau ^2)\\
&\bar{d}(y,s)=d(x_0+y\tau ,s\tau ^2).
\end{split}
\eeq
Then $(\bar{u},\bar{P},\bar{d})$ is a weak solution to
(\ref{lce}) on $P_1(0,1)$, and by (\ref{u-lcf-3}),
\beq\label{u-lcf-4} \|\bar u\|_{L^p_t
L^q_x(P_1(0,1))} +\|\nabla \bar d\|_{L^p_t L^q_x(P_1(0,1))}\leq
\epsilon.
\eeq
By Lemma \ref{unqlc-lemma3.1}, we conclude that
\beq\label{u-lcf-5}
|\bar u(0,1)|+|\nabla \bar d(0,1)|\leq C
\left(\|\bar u\|_{L^p_t L^q_x(P_1(0,1))} +\|\nabla \bar d\|_{L^p_t
L^q_x(P_1(0,1))}\right).
\eeq
By rescaling, this implies
\beq\label{u-lcf-6}
\tau\left(| u(x_0,\tau^2)|+|\nabla d(x_0,\tau^2)|\right)\leq C
\left(\|u\|_{L^p_t L^q_x(\mathbb R^n\times[0,\tau^2])}
+\|\nabla d\|_{L^p_t L^q_x(\mathbb R^n\times[0,\tau^2])}\right)\leq C\epsilon.
\eeq
Taking supremum over all $x_0\in\mathbb R^n$ completes the proof.
\endpf
\medskip
\noindent{\bf Proof of Theorem \ref{unique3}}: By (\ref{unqlc3.36}), we have that for any $\epsilon>0$,
there exists $t_0=t_0(\epsilon)>0$ such that
\begin{eqnarray}\label{max_bound}
\mathcal {A}(t_0)&=&\sum_{i=1}^2\Big[\sup_{0\le t\le t_0}\sqrt{t}(\|u_i(t)\|_{L^{\infty}(\R^n)}+\|\nabla
d_i(t)\|_{L^{\infty}(\R^n)})\nonumber\\
&&+(\|u_i\|_{L^p_t L^{q}_x(\R^n\times [0,t_0]))}+\|\nabla
d_i\|_{L^p_t L^{q}_x(\R^n\times [0,t_0]))})\Big]\le \epsilon.
\end{eqnarray}
It suffices to show $(u_1,d_1)=(u_2,d_2)$ on $\mathbb R^n\times [0,t_0]$. To do so,
let $u=u_1-u_2$ and $d=d_1-d_2$.
Applying $\mathbb{P}$ to both (\ref{lce})$_1$ for $u_1$ and $u_2$ and taking the difference of
resulting equations, we have that
\begin{equation}
\begin{cases}
u_t-\mathbb{A} u=-\mathbb{P}\nabla\cdot\left(u\otimes u_1+u_2\otimes u
+\nabla d\otimes\nabla d_1+\nabla d_2\otimes\nabla d+(|\nabla d_1|+|\nabla d_2|)|\nabla d|\mathbb I_n\right), \\
\nabla\cdot u=0, \\
d_t-\Delta d=[(\nabla d_1+\nabla d_2)\cdot\nabla d\ d_2+|\nabla d_1|^2d]
-[u\cdot\nabla d_1+u_2\cdot\nabla d], \\
(u,d)\Big|_{t=0}=(0,0).
\end{cases}
\end{equation}
By the Duhamel formula, we have that for any $0<t\le t_0$,
\beq\nota
\begin{split}
u(t)=-\int_0^te^{-(t-\tau)\mathbb{A}}\mathbb{P}\nabla\cdot\Big(u\otimes u_1+u_2\otimes u
+\nabla d\otimes\nabla d_1+\nabla d_2\otimes\nabla d+(|\nabla d_1|+|\nabla d_2|)|\nabla d|\mathbb I_n\Big)\,d\tau,
\end{split}
\eeq
\beq\label{unqlc4.6}
\begin{split}
d(t)=\int_0^te^{-(t-\tau)\Delta}\Big((\nabla d_1+\nabla d_2)\cdot\nabla
d \ d_2+|\nabla d_1|^2d -u\cdot\nabla d_1-u_2\cdot\nabla
d\Big)\,d\tau.
\end{split}
\eeq
For $0<t\le t_0$, set
$$\Phi(t)=\|u\|_{L^p_tL^{q}_x(\R^n\times [0,t]))}+\|\nabla
d\|_{L^p_tL^{q}_x(\R^n\times [0,t]))}+\sup\limits_{0\leq\tau\leq t}\|d(\cdot,\tau)\|_{L^{\infty}(\R^n)}.$$
By (\ref{unqlc4.6}) and the standard estimate on the heat kernel, we obtain that
\beq\label{unqlc4.9}
\begin{split}
\Big\|\nabla
d(t)\Big\|_{L^{q}(\R^n)}\leq &C\Big[\sum_{i=1}^2\int_0^{t}(t-\tau)^{\frac{1}{p}-1}
\|\nabla d_i\|_{L^{q}(\R^n)}\|\nabla
d\|_{L^{q}(\R^n)}\,d\tau\\
&+\|d\|_{L^{\infty}(\R^n)}\int_0^{t}(t-\tau)^{\frac{1}{p}-1}
\|\nabla d_1\|_{L^{q}(\R^n)}^2\,d\tau\\
&+\int_0^{t}(t-\tau)^{\frac{1}{p}-1}
\|\nabla d_1\|_{L^{q}(\R^n)}\|u\|_{L^{q}(\R^n)}\,d\tau\\
&+\int_0^{t}(t-\tau)^{\frac{1}{p}-1}
\|u_2\|_{L^{q}(\R^n)}\|\nabla d\|_{L^{q}(\R^n)}\,d\tau\Big].
\end{split}
\eeq
By the standard Riesz potential estimate in $L^p$-spaces (see \cite{FJR} Theorem 3.0), we see that
$\nabla d\in L^p_tL^q_x(\mathbb R^n\times [0,t_0])$, and
\beq\label{unqlc4.10}
\begin{split}
\Big\|\nabla d\Big\|_{L^p_tL^q_x(\R^n\times [0,t_0])}\leq&C\Big[\sum_{i=1}^2\|\nabla
d_i\|_{L^p_tL^{q}_x(\R^n\times [0,t_0]))}\|\nabla
d\|_{L^p_tL^{q}_x(\R^n\times [0,t_0])}\\
&+\|d\|_{L^{\infty}(\R^n\times[0,t_0])}\|\nabla
d_1\|^2_{L^p_tL^{q}_x(\R^n\times [0,t_0])}\\
&+\|\nabla
d_1\|_{L^p_tL^{q}_x(\R^n\times [0,t_0])}\|u\|_{L^p_tL^{q}_x(\R^n\times [0,t_0])}\\
&+\|u_2\|_{L^p_tL^{q}_x(\R^n\times [0,t_0])}\|\nabla
d\|_{L^p_tL^{q}_x(\R^n\times [0,t_0])}\Big]\\
\leq&C\mathcal A(t_0)\Phi(t_0).
\end{split}
\eeq
Similarly, by using the estimate of Theorem 3.1 (i) of \cite{FJR}, we have that
$u\in L^p_tL^q_x(\mathbb R^n\times [0,t_0])$, and
\beq\label{unqlc4.11}
\begin{split}
\|u\|_{L^p_tL^{q}_x(\R^n\times [0,t_0])} \leq& C\mathcal A(t_0)\Phi(t_0).
\end{split}
\eeq
Now we need to estimate
$\sup\limits_{0\leq\tau\leq t_0}\|d(\cdot,\tau)\|_{L^{\infty}(\R^n)}$. We claim
\begin{equation}\label{max_estimate_d}
\|d\|_{L^\infty(\mathbb R^n\times [0,t_0])}\le C\mathcal A(t_0)\Phi(t_0).
\end{equation}
To show (\ref{max_estimate_d}), let
$\displaystyle H(x,t)$
be the heat kernel of $\R^n$.
By (\ref{unqlc4.6}), we have
\beq\label{unqlc3.35}
\begin{split}|d(x,t)|=&\Big|\int_0^t\int_{\R^n}H(x-y,t-\tau)\left((\nabla d_1+\nabla d_2)\cdot\nabla d\ d_2+|\nabla
d_1|^2d\right)(y,\tau)\,dyd\tau\\
&-\int_0^t\int_{\R^n}H(x-y,t-\tau)\left(u\cdot\nabla d_1+u_2\cdot\nabla d\right)(y,\tau)\,dyd\tau\Big|\\
\leq& C\Big[\int_0^t\int_{\mathbb R^n} H(x-y,t-\tau)K(y,\tau)\,dyd\tau\\
&\quad +\int_0^t\int_{\R^n} H(x-y,t-\tau)|\nabla d_1|^2(y,\tau)\,dyd\tau
\cdot\sup\limits_{0\leq\tau\leq
t}\|d(\cdot,\tau)\|_{L^{\infty}(\R^n)}\Big], \\
\end{split}\eeq
where
$$K(y,\tau):=\sum\limits_{i=1}^2(|u_i|+|\nabla d_i|)(|u|+|\nabla d|)(y,\tau).$$
By (\ref{max_bound}), we have that for any $0<t\le t_0$,
\beq\label{unqlc3.37}
\begin{split}
&\int_0^t\int_{\mathbb R^n} H(x-y,t-\tau)K(y,\tau)\,dyd\tau\\
\leq&\mathcal{A}(t_0)\int_0^{t}(t-\tau)^{-\frac{n}2}\tau^{-\frac12}\int_{\mathbb R^n}(|u|+|\nabla d|)\exp\Big(-\frac{|x-y|^2}{4(t-\tau)}\Big)
\,dyd\tau\\
\leq&\mathcal{A}(t_0)\Big\|(t-\tau)^{-\frac{n}{2q}}\tau^{-\frac12}\Big\|_{L^{\frac{p}{p-1}}([0,t])}
\Big\||u|+|\nabla d|\Big\|_{L^p_tL^q_x(\mathbb R^n\times [0,t])}\\
\leq& C\mathcal {A}(t_0)\Phi(t_0),
\end{split}\eeq
where we have used H\"older inequality and
\begin{eqnarray*}
\Big\|(t-\tau)^{-\frac{n}{2q}}\tau^{-\frac12}\Big\|_{L^{\frac{p}{p-1}}([0,t])}^{\frac{p}{p-1}}
&=&t^{(\frac12-(\frac{n}{2q}+\frac{1}{p}))\frac{p}{p-1}}
\int_0^1 (1-\tau)^{-\frac{np}{2(p-1)q}}\tau^{-\frac{p}{2(p-1)}}\,d\tau\\
&=&\int_0^1 (1-\tau)^{-\frac{p-2}{2(p-1)}}\tau^{-\frac{p}{2(p-1)}}\,d\tau<+\infty,
\end{eqnarray*}
as (i) $\frac{n}{2q}+\frac{1}{p}=\frac12$,
and (ii) $2<p<+\infty$ yields $\frac{p}{2(p-1)}<1$ and $\frac{p-2}{2(p-1)}<1$.
Similarly, we can obtain that for $0\le t\le t_0$,
\begin{equation}\label{unqlc3.38}
\int_0^t\int_{\R^n} H(x-y,t-\tau)|\nabla d_1|^2(y,\tau)\,dyd\tau
\le C\mathcal {A}^2(t_0).
\end{equation}
Putting (\ref{unqlc3.37}) and (\ref{unqlc3.38}) into (\ref{unqlc3.35}) and taking supremum over $(x,t)\in\mathbb R^n\times [0,t_0]$, we have
\begin{equation}\label{max_bound1}
\sup_{0\le t\le t_0}\|d\|_{L^\infty(\mathbb R^n)}
\le C\mathcal{A}(t_0)\Phi(t_0)+C\mathcal {A}^2(t_0)\sup_{0\le t\le t_0}\|d\|_{L^\infty(\mathbb R^n)}.
\end{equation}
Therefore, if we choose $\epsilon\le \sqrt{\frac{1}{2C}}$ so that
$C\mathcal{A}^2(t_0)\le C\epsilon^2\le \frac12$, then we obtain (\ref{max_estimate_d}).
Putting (\ref{unqlc4.10}), (\ref{unqlc4.11}) and (\ref{max_estimate_d}) together, and choosing $\epsilon\le\frac{1}{2C}$,
we obtain
$$\Phi(t_0)\leq C\mathcal{A}(t_0)\Phi(t_0)\le \frac12\Phi(t_0).$$
This implies that $\Phi(t_0)=0$ and hence $(u_1,d_1)\equiv (u_2,d_2)$ on $\mathbb R^n\times [0,t_0]$.
If $t_0<T$, then we can repeat the argument for $t\in [t_0,T]$ and eventually show that $(u_1,d_1)\equiv (u_2,d_2)$ on $\mathbb R^n\times [0,T]$.
This completes the proof. \endpf
\vspace{4mm}
\noindent{\bf Acknowledgements}. The paper is part of my Ph.D. thesis in University of Kentucky. I would like to thank my advisor Professor Changyou Wang for his helpful discussion and constant encouragement.
|
\section{Introduction}\label{sec:intro}
The requirement that preferences are \emph{coherent} aims to make rigorous
the idea that elementary restrictions on rational preferences entail
that personal probabilities satisfy the axioms of mathematical
probability. This use of coherence as a justification of personal
probability is very well illustrated by de Finetti's (\citeyear{deFinetti1974}) approach to the
foundations of probability. De Finetti distinguished two senses
of coherence: coherence$_1$ and coherence$_2$. Coherence$_1$ requires
that probabilistic forecasts for random variables (he calls them
previsions) do not lead to
a finite set of fair contracts that, together, are uniformly dominated
by abstaining. Coherence$_2$ requires that a finite set of probabilistic
forecasts cannot be uniformly dominated under Brier (squared error)
score by a rival set of forecasts. He showed that these two senses of
coherence are equivalent in the following sense. Each version of
coherence results in using the expectation of a random variable as its
forecast. Moreover, these expectations are
based on a finitely additive probability without requiring that
personal probability is countably additive. [In Appendix \ref{app:lf},
we explain what we mean by expectations with respect to finitely
additive probabilities. These are similar in many ways, but not
identical to integrals in the sense of \citet{dunford-schwartz1958},
Chapter III.] Schervish, Seidenfeld and
Kadane (\citeyear{ssk2009b})
extended this equivalence to include a large class of strictly proper
scoring rules (not just Brier score) but for events only. The
corresponding extension to general random variables is included in the
supplemental article [Schervish, Seidenfeld and Kadane
(\citeyear{ssk2013})]. Here, we refer to the
extended sense of coherence$_2$ as coherence$_3$.
We investigate asymmetries between coherence$_1$ and coherence$_3$
reflecting differences between cases where personal probabilities are
countably additive and where personal probabilities are finitely (but
not countably) additive. We give conditions where coherence$_3$ may be
applied to assessing countably many forecasts at once, but where
coherence$_1$ cannot be applied to combining infinitely many fair
contracts. Also, we study conditional forecasts given elements of a
partition $\pi$, where the conditional forecasts are based on the
conditional probabilities
given elements of $\pi$. Each coherence criterion is violated by combining
infinitely many conditional forecasts when those
conditional forecasts are not conglomerable (see
Definition~\ref{def:nonc}) in the partition
$\pi$. Neither criterion is violated by combining infinitely many
conditional forecasts when conditional expectations satisfy the law of
total previsions (see Definition~\ref{def:ltp}) in $\pi$.
\section{Results of de Finetti}\label{sec:results} Coherence of
preference, as de Finetti [(\citeyear{deFinetti1974}), Chapter~3] formulates it, is the
criterion that a rational decision maker respects \textit{uniform
(strict) dominance}. In Section~\ref{sub:dominance}, we
explain the version of the Dominance Principle that de Finetti uses.
In Section~\ref{sub:coherence}, we review de Finetti's two versions of
coherence, with a focus on how preferences based on a finitely
additive probability are coherent.
\subsection{Dominance}\label{sub:dominance} Let $\Omega$ be a
set. The elements of $\Omega$ will be
called \emph{states} and denoted $\omega$. Random variables are
real-valued functions with domain $\Omega$, which we denote with
capital letters. Let $I$ index a set of options. Consider a
hypothetical decision problem ${\cal O}$ specified by a set of
exclusive options ${\cal O} = \{O_i\dvtx i \in I\}$. Each option $O_i$
is a random variable with the following interpretation: If $\omega$
is the state which occurs, then $O_i(\omega)$ denotes the decision
maker's loss (negative of cardinal utility) for choosing option
$O_i$. The values of $O_{i}$ (for all
$i\in I$) are defined up to a common positive affine transformation.
\begin{definition} Let $O_i$ and $O_j$
be two options from ${\cal O}$. If there exists an $\varepsilon> 0$
such that for each $\omega\in\Omega$, $O_j(\omega) > O_i(\omega) +
\varepsilon$, then option
$O_i$ \emph{uniformly strictly dominates $O_j$}.
If, for each $\omega$, $O_j(\omega) > O_i(\omega)$, we say that $O_i$
\emph{simply dominates}~$O_j$.
\end{definition}
Uniform strict dominance is clearly stricter than simple dominance.
As we explain, next, in order to permit preferences based on
maximizing finitely (and not necessarily countably) additive
expectations, de Finetti used the following Dominance Principle,
rather than
some other more familiar concepts of admissibility, for example, simple
dominance. There are additional ways to define dominance, which
we discuss further in Section~\ref{sec:summary}.
\textsc{Dominance Principle}: Let $O_i$ and $O_j$ be options in
${\cal O} $. If $O_i$ uniformly (strictly) dominates $O_j$, then
$O_j$ is an \textit{inadmissible} choice from ${\cal O}$.
\subsection{Coherence$_1$ and coherence$_2$}\label{sub:coherence}
De Finetti [(\citeyear{deFinetti1974}), Chapter~3] formulated two criteria of
\textit{coherence} that are based on the Dominance Principle.
Throughout this paper, we follow the convention of identifying events
with their indicator functions.\vspace*{-2pt}
\begin{definition}\label{def:coh}\label{def:condcoh}
A \emph{conditional prevision} (or \emph{conditional forecast})
$P(X|H)$ for a
random variable $X$ given a nonempty event $H$ is a fair price for
buying and selling $X$ in the
sense that, for all real $\alpha$, the option that costs the agent
$\alpha H[X-P(X|H)]$ is considered fair. [We call $P(X|\Omega)$ an
\emph{unconditional prevision} and denote it $P(X)$.]
A collection $\{P(X_i|H_i)\dvtx i\in
I\}$ of such conditional forecasts is \emph{coherent$_1$} if, for
every finite
subset $\{i_1,\ldots,i_n\}\subseteq I$ and all real
$\alpha_1,\ldots,\alpha_n$, there exists no $\varepsilon>0$ such that
\[
\sum_{j=1}^n\alpha_jH_{i_j}(
\omega)\bigl[X_{i_j}(\omega)-P(X_{i_j}|H_{i_j})\bigr]
\geq\varepsilon
\]
for all $\omega\in\Omega$.
A collection of conditional forecasts is \emph{coherent$_2$} if
no sum of finitely many (Brier score) penalties
can be uniformly strictly dominated in the partition of states by the
sum of penalties from a rival set of forecasts for the same random
variables. That is, for every finite subset
$\{i_1,\ldots,i_n\}\subseteq I$, all alternative forecasts
$q_{i_1},\ldots,q_{i_n}$, and all positive $\alpha_1,\ldots,\alpha_n$,
there is no $\varepsilon>0$ such that\looseness=-1
\[
\sum_{j=1}^n\alpha_jH_{i_j}(
\omega)\bigl[X_{i_j}(\omega)-P(X_{i_j}|H_{i_j})
\bigr]^2 \geq\sum_{j=1}^n
\alpha_jH_{i_j}(\omega)\bigl[X_{i_j}(\omega
)-q_{i_j}\bigr]^2+\varepsilon
\]\looseness=0
for all $\omega$.
\end{definition}
De Finetti [(\citeyear{deFinetti1974}), pages~88--89] proved that a decision maker who wishes
to be both coherent$_1$ and coherent$_2$ must choose the same
forecasts for both purposes.
He also proved that the decision maker's coherent$_1$ forecasts are
represented by a finitely additive personal
probability, $P(\cdot)$, in the sense of Definition~\ref{def:fap} below.
If $P(H)=0$, then coherence$_1$ and coherence$_2$ place no
restrictions on $P(X|H)$ for bounded $X$.
Nevertheless, it is possible and useful to make certain intuitive
assumptions about conditional forecasts given events with 0
probability. In particular, Theorems \ref{thm:3} and \ref{thm:2008b}
of this paper assume that $P(\cdot|H)$ is a
finitely additive expectation (in the sense of
Definition~\ref{def:daniell} in Appendix \ref{app:lf}) satisfying
$P(X|H)=P(HX|H)$ for all $H$ and $X$. This assumption holds whenever
$P(H)>0$, and it captures the idea that $P(\cdot|H)$ is concentrated
on $H$.
De Finetti [(\citeyear{deFinetti1975}), Appendix~16] introduces an axiom that places a
similar requirement on conditional previsions. See Levi
(\citeyear{Levi1980}), Section~5.6, and \citet{regazzini1987} for other ways
to augment the coherence criteria of Definition~\ref{def:condcoh}
in
order to satisfy these
added requirements on conditional previsions given a null event.
Rather than adding such requirements to the definition of coherence, we\vadjust{\goodbreak}
prefer that individual agents who wish to adopt them do so as
explicit additional assumptions. Example~2 in the
supplemental article [Schervish, Seidenfeld and
Kadane (\citeyear{ssk2013})] illustrates our reason
for such a preference. In this way, our definition of coherence
is slightly weaker than that of de Finetti.
As an aside, the meaning of conditional expected value in the
finitely-additive theory differs from its meaning in the
countably-additive theory in this one major regard: In the
finitely-additive theory a conditional expectation can be specified given
an arbitrary nonempty event, regardless of whether that event has positive
probability. A conditional expectation of a bounded random variable
given an event with zero
probability is not defined uniquely in terms of unconditional
expectations, but \citet{dubins1975} shows that, in the finitely additive
theory, conditional expectations can be defined on the set of bounded
random variables so that they are
finitely additive expectations. In the countably-additive theory,
conditional expectation is defined twice: given events with positive
probability and given $\sigma$-fields. The two definitions match in a
well-defined way, and both provide uniquely defined conditional
expectations in
terms of unconditional expectations.
\begin{definition}\label{def:fap}
A probability $P(\cdot)$ is \emph{finitely additive} provided that,
when events $F$ and $G$ are disjoint, that is, when $F \cap G =
\varnothing$,
then $P(F \cup G) = P(F) + P(G)$.
A probability is \emph{countably additive} provided that when $F_{i}\ (i = 1, \ldots)$ is a denumerable sequence of pairwise disjoint
events, that is, when $F_{i} \cap F_{j} = \varnothing$ if $i \neq j$, then
$P(\bigcup_{i=1}^\infty F_{i}) = \sum_{i=1}^\infty P(F_{i})$.
We call a probability $P$ \emph{merely finitely additive} when $P$
is finitely but not countably additive. Likewise, then its
$P$-expectations are merely finitely additive.
\end{definition}
For each pair $X$ and $Y$ of random variables with finite previsions
(expectations), $P(X +Y) = P(X) + P(Y)$.
For countably additive expectations and countably many random
variables $\{X_{i}\}_{i=1}^\infty$, conditions under which
$P (\sum_{i=1}^\infty X_i ) = \sum_{i=1}^\infty P(X_i)$ can
be derived from various theorems such as the monotone convergence theorem,
the dominated convergence theorem, Fubini's theorem and Tonelli's theorem.
De Finetti (\citeyear{deFinetti1981}) recognized that coherence$_2$ (but not
coherence$_1$) provided an incentive compatible solution to the
problem of mechanism design for eliciting a coherent set of personal
probabilities. Specifically, Brier score is a \textit{strictly
proper scoring rule}, as defined here.
\begin{definition}\label{def:4} A \emph{scoring rule} for coherent
forecasts of a random
variable $X$ is a real-valued loss function $g$ with two real arguments:
a value of the random variable and a forecast $q$.
Let $\P_g$ be the collection of probability distributions such that
$P(X)$ is finite and $P[g(X,q)]$ is finite for at least one $q$.
We say that $g$ is \emph{proper} if, for every probability $P\in\P_g$,
$P[g(X,q)]$ is minimized (as a function of $q$) by $q=P(X)$.
If, in addition, only the quantity $q=P(X)$ minimizes expected score,
then the scoring rule is \emph{strictly proper}.\vadjust{\goodbreak}
\end{definition}
The following trivial result connects proper scoring rules with
conditional distributions.
\begin{proposition}\label{pro:src}
If $H$ is a nonempty event and $P(\cdot|H)$ is a probability distribution
then $P[g(X,q)|H]$ is (uniquely) minimized by $q=P(X|H)$ if $g$ is
(strictly) proper.
\end{proposition}
Some authors reserve the qualification strictly proper for scoring
rules that are designed to elicit an entire distribution, rather than
just the mean of a distribution.
[See Gneiting (\citeyear{gneiting11a}), who calls the
latter kind \emph{strictly consistent}.] For the remainder of this
paper, we
follow the language of Definition~\ref{def:4}, which matches the usage in
Gneiting (\citeyear{gneiting11b}).
We present some background on strictly proper scoring rules in
Section~\ref{sec:scoring}. Section~\ref{sec:extensions} gives our
main results.
We discuss propriety of scoring rules for infinitely many forecasts in
Section~\ref{sec:proper}.
\section{Background on strictly proper scoring rules}\label{sec:scoring}
In this section, we introduce a large class of strictly proper scoring
rules that we use as generalizations of Brier score. Associated with
this class, we introduce a third coherence concept that generalizes
coherence$_2$.
\begin{definition}
Let ${\cal C}$ be a class of strictly proper scoring rules. Let
$\{(X_i, H_i)\dvtx i\in I\}$ be a collection of random variable/nonempty
event pairs with corresponding conditional forecasts $\{p_i\dvtx i\in I\}$.
The forecasts are\break \emph{coherent$_3$ relative to ${\cal C}$} if, for
every finite subset $\{i_{j}\dvtx j=1,\ldots,n\}\subseteq I$, every set of scoring rules
$\{g_j\}_{j=1}^n\subseteq{\cal C}$, and every set $\{q_j\}_{j=1}^n$ of
alternative forecasts, there is no $\varepsilon>0$ such that
\[
\sum_{j=1}^nH_{i_j}(
\omega)g_j\bigl(X_{i_j}(\omega),p_{i_j}\bigr)\geq
\sum_{j=1}^n H_{i_j}(
\omega)g_j\bigl(X_{i_j}(\omega),q_j\bigr)+
\varepsilon
\]
for all $\omega$.
That is, no sum of finitely many scores can be uniformly strictly
dominated by the sum of scores from rival forecasts.
\end{definition}
Coherence$_2$ is the special case of coherence$_3$ in which ${\cal C}$
consists solely of Brier score. The supplemental article
[Schervish, Seidenfeld and
Kadane (\citeyear{ssk2013})] includes a proof
that, if ${\cal C}$ consists of strictly proper
scoring rules of the form (\ref{eq:gdef}) below, then coherence$_3$
relative to ${\cal C}$ is equivalent to coherence$_1$.
The general form of scoring rule that we will consider is
\begin{equation}
\label{eq:gdef} g(x,q)=\cases{
\displaystyle\int
_x^q(v-x)\,d\lambda(v),&\quad $\mbox{if $x\leq q$,}$
\vspace*{2pt}\cr
\displaystyle\int_q^x(x-v)\,d\lambda(v),&\quad $\mbox{if $x>q$,}$}
\end{equation}
where $\lambda$ is a measure that is mutually absolutely continuous with
Lebesgue measure and is finite on every bounded interval.
It is helpful to rewrite (\ref{eq:gdef}) as
\begin{equation}
\label{eq:gsimp} g(x,q)=\int_q^x(x-v)\,d\lambda(v),
\end{equation}
using the convention that an integral whose limits are in the wrong
order equals the negative of the integral with the limits in the
correct order. Another interesting way to rewrite (\ref{eq:gdef}),
using the same convention, is
\begin{equation}
\label{eq:glin} g(x,q)=\lambda\bigl((q,x)\bigr)\bigl[x-r(x,q,\lambda)\bigr],
\end{equation}
where, for all $a$ and $b$,
\begin{equation}
\label{eq:ra2} r(a,b,\lambda)=\frac{\int_a^bv\,d\lambda(v)}{\lambda((a,b))}.
\end{equation}
An immediate consequence of (\ref{eq:glin}) is that, if $p$ and $q$
are real numbers, then
\begin{equation}
\label{eq:diff} g(x,q)-g(x,p)=\lambda\bigl((q,p)\bigr)\bigl[x-r(q,p,\lambda)
\bigr].
\end{equation}
The form (\ref{eq:gdef}) is suggested by equation (4.3) of \citet{savage71}.
Each such scoring rule is finite, nonnegative and continuous as a
function of $(x,q)$. If we wanted to
consider only countably additive distributions, we could use a larger
class of scoring rules by allowing $\lambda$ to be an infinite measure
supported on a bounded interval $(c_1,c_2)$. But this relaxation would
allow functions $g$ that are not strictly proper for natural classes
of finitely additive distributions. Example~\ref{exa:log} below
illustrates this point. Lemma~\ref{lem:sp} justifies the
use of (\ref{eq:gdef}) as the form of our scoring rules. The proofs of
all results in the body of the paper are given in Appendix \ref{app:aux}.
\begin{lemma}\label{lem:sp}
Let $g$ be a scoring rule of the form (\ref{eq:gdef}). Then $g$ is strictly
proper.
\end{lemma}
It follows from (\ref{eq:glin}) that, if $\lambda$ is a probability
measure with finite mean, then $\P_g$ from Definition~\ref{def:4} is
the class of all finitely additive distributions with finite mean because
$\lambda((q,x))$ and $\lambda((q,x))r(q,x,\lambda)$ are both bounded
functions of $q$ and $x$. Even if $\lambda$ is not a finite measure,
(\ref{eq:diff}) implies that, if $P[g(X,q_0)]$ is finite, then
$h(x,p)=g(x,p)-g(x,q_0)$ is linear in $x$ so that $h(x,p)$ is also
strictly proper with $\P_h$ equal to the class of all probabilities
with finite mean. For example, if $g(x,p)=(x-p)^2$, namely Brier score, then
$\P_g$ is the set of distributions with finite second moment. However,
$h(x,p)=(x-p)^2-x^2$ has $\P_h$ equal to the class of all probabilities
with finite mean.
Let $f(\cdot)$ denote the Radon--Nikodym derivative of
$\lambda$ with respect to Lebes\-gue measure.
Some familiar examples of strictly proper scoring rules are recovered
by setting $f$ equal to specific functions. Brier score corresponds
to $f(v)\equiv2$. Logarithmic score on the interval $(c_1,c_2)$ corresponds
to $f(v)=(c_2-c_1)/[(c_2-v)(v-c_1)]$, but the corresponding measure is
infinite on $(c_1,c_2)$. Hence, logarithmic score is not of the form
(\ref{eq:gdef}).
In addition, if $g$ is this logarithmic score, then $\P_g$ does not
include all finitely additive distributions that take values in the
bounded interval $(c_1,c_2)$, as the following example illustrates.
\begin{example}\label{exa:log}
Let $X$ be a random variable whose entire distribution
is agglutinated at $c_1$ from above. That is, let $P(X>c_1)=1$ and
$P(X<c_1+\varepsilon)=1$ for all $\varepsilon>0$. Let $g$ be the
logarithmic scoring rule that uses $f(v)$ from above.
Then $P(X)=c_1$, but $g(X(\omega),c_1)=\infty$ for all $\omega$,
which could not have finite mean even if we tried to extend the definition
of random variables to allow them to assume infinite values. On the
other hand, for $c_1<q<c_2$, the mean of $g(X,q)$ is
$\log[(c_2-c_1)/(c_2-q)]>0$, which decreases to 0 as $q$ decreases to
$c_1$, and is always finite. So, $\P_g$ is nonempty but does not
contain~$P$.
\end{example}
Some of our results rely on one or another condition that prevents the
$\lambda$ measures that determine the scoring rules from either being
too heavily concentrated on small sets or from being too different
from each other.
\begin{definition}
Let ${\cal C}=\{g_i\dvtx i\in I\}$ be a collection of strictly proper
scoring rules of the form (\ref{eq:gdef}) with corresponding measures
$\{\lambda_i\dvtx i\in I\}$.
\begin{longlist}[(ii)]
\item[(i)] Suppose that, for every $\varepsilon\geq0$, there exists
$\delta_\varepsilon>0$ such that for all $i\in I$ and all real $a<b$,
$\lambda_i((a,b))>\varepsilon$ implies $a+\delta_\varepsilon\leq
r(a,b,\lambda
_i)\leq
b-\delta_\varepsilon$.
Then we say that the collection ${\cal C}$ satisfies the \emph{uniform
spread condition}.
\item[(ii)] Suppose that, for every $\varepsilon>0$ and every $i\in I$,
there exists $\gamma_{i,\varepsilon}>0$ such that for all $j\in I$ and
all real $a<b$, $\lambda_i((a,b))\geq\varepsilon$ implies
$\lambda_j((a,b))\geq\gamma_{i,\varepsilon}$. Then
we say that the collection ${\cal C}$ satisfies the \emph{uniform
similarity condition}.
\end{longlist}
\end{definition}
The $r(a,b,\lambda)$ in (\ref{eq:ra2}) can be thought of as the mean
of the probability measure on the interval $(a,b)$ obtained by
normalizing $\lambda$ on the interval. The uniform spread condition
insures that, the $\lambda$ measures are spread out enough to keep the means
of the normalized measures on intervals far enough away from both endpoints.
The next result gives sufficient conditions for both the uniform similarity
and uniform spread conditions. It is easy to see that the conditions
are logically independent of each other.
\begin{lemma}\label{lem:ubf}
Let ${\cal C}=\{g_i\dvtx i\in I\}$ be a collection of strictly proper
scoring rules of the form (\ref{eq:gdef}) with corresponding measures
$\{\lambda_i\dvtx i\in I\}$ and corresponding Radon--Nikodym derivatives
$\{f_i\dvtx i\in i\}$ with respect to Lebesgue measure.
\begin{longlist}[(ii)]
\item[(i)] Assume that there exists $U<\infty$ such
that $f_i(v)\leq U$, for all $v$ and all $i\in I$. Then ${\cal C}$
satisfies the uniform spread condition.
\item[(ii)] Assume that for every $i\in I$, there exists $L_i>0$ such
that $f_j(v)/f_i(v)\geq L_i$, for all $v$ and all $j\in I$. Then ${\cal C}$
satisfies the uniform similarity condition.
\end{longlist}
\end{lemma}
As an example, suppose that each
$\lambda_i$ is $\alpha_i>0$ times Lebesgue measure. If the
$\alpha_i$ are bounded above, then ${\cal C}$
satisfies the uniform spread condition. If the $\alpha_i$ are bounded
away from 0, then ${\cal C}$ satisfies the uniform similarity
condition. These sets of measures correspond to
multiples of Brier score. There are collections that satisfy the uniform
spread condition without satisfying the conditions of part (i) of
Lemma~\ref{lem:ubf}. For example, let $f(v)=|v|^{-1/2}/2$ which is
not bounded above. For this~$f$, we have
$\lambda((a,b))=|\sqrt{|b|}-\sqrt{|a|}|$ if $0\notin(a,b)$, and
$\lambda((a,b))=\sqrt{|a|}+\sqrt{|b|}$ if $0\in(a,b)$.
So $\lambda((a,b))^2$ is no larger than two times the distance between
$a$ and~$b$. Also, $r(a,b,\lambda)$ is always at least $1/3$ of the way
from both $a$ and $b$. We can add the corresponding scoring rule to
any class that already satisfies the uniform spread condition by (if
necessary) lowering $\delta_\varepsilon$ to $\varepsilon^2/6$.
\section{Extensions to countably many options}\label{sec:extensions}
In Section~\ref{sub:dominance3}, we investigate when each sense of
coherence can be extended to allow combining countably many forecasts into
a single act by summing together their individual outcomes. In
Section~\ref{sub:dominance4}, we introduce the concept of \emph{conditional} forecasts and present results about the
combination of countably many coherent conditional forecasts.
\subsection{Dominance for countably many forecasts}\label{sub:dominance3}
Let $\{X_i\}_{i=1}^\infty$ be a countable set of random variables
with corresponding coherent$_1$ unconditional previsions $\{p_i\}
_{i=1}^\infty$.
Let $\{\alpha_i\}_{i=1}^\infty$ be a sequence of real numbers.
The decision maker's net loss
in state $\omega$, from adding the individual losses from the
fair options $\alpha_i[X_i(\omega)-p_i]$ is
\begin{equation}
\label{eq:infsumprev} \sum_{i=1}^\infty
\alpha_{i}\bigl[X_{i}(\omega) - p_{i}\bigr].
\end{equation}
Similarly, if the agent's prevision $p_i$ for $X_i$ is scored by the
strictly proper scoring rule $g_i$ for each $i$, the total score in
each state $\omega$ equals
\[
\sum_{i=1}^\infty g_i
\bigl(X_i(\omega), p_{i}\bigr).
\]
We assume that each of the two series above are convergent for all
$\omega\in\Omega$.
\begin{example}[(Combining countably many
forecasts)]\label{examp:1} De
Finetti [(\citeyear{deFinetti1949}), page~91] noted that when the decision
maker's personal probability is merely finitely additive, she/he
cannot always accept as fair the countable sum~(\ref{eq:infsumprev})
determined by coherent$_{1}$ forecasts. That sum may be uniformly
dominated by abstaining. Let $\Omega=
\{\omega_i\}_{i=1}^\infty$ be a countable state space. Let $W_i$ be
the indicator function for state $\omega_i\dvtx W_i(\omega) = 1$ if
$\omega
= \omega_i$ and $W_i(\omega) = 0$ if $\omega\neq\omega_i$.
Consider a collection of merely finitely additive coherent$_1$ forecasts
$P(W_i) = p_i \geq0$ where $\sum_{i=1}^\infty p_{i} = c < 1$. So
$P(\cdot)$ is not countably additive. With
$\alpha_i=1$, for all $i$, the loss from combining these
infinitely many forecasts into a single option is uniformly positive,
\[
\sum_{i=1}^\infty\alpha_i
\bigl[W_i(\omega) - p_i\bigr] =(1-c)>0.
\]
Hence, the
decision maker's alternative to abstain, with constant loss $0$,
uniformly strictly dominates this infinite combination of fair options.
If, on the other hand, the decision maker's personal probability $P$
is countably additive, then $c=1$. For arbitrary
$\{\alpha_i\}_{i=1}^\infty$ such that $d=\sum_{i=1}^\infty\alpha_ip_i$
is defined and finite, the sum of losses is
\begin{equation}
\label{eq:gca} \sum_{i=1}^\infty
\alpha_i\bigl[W_i(\omega) - p_i\bigr] =
\alpha_{i(\omega)}-d,
\end{equation}
where $i(\omega)$ is the unique $i$ such that $W_i(\omega)=1$.
Because $c=1$, there is at least one $\alpha_i\leq d$ and at least one
$\alpha_i\geq d$, hence (\ref{eq:gca}) must be nonpositive for at
least one~$i$, and abstaining does not uniformly strictly dominate.
\end{example}
Next, we focus on the parallel question whether a coherent$_3$ set
of forecasts remains undominated when strictly proper scores for
countably many forecasts are summed together. Some conditions will
be needed in order to avoid $\infty-\infty$ arising in the
calculations, and these are stated precisely in the theorems.
The principal difference between dominance for infinite sums of
forecasts and dominance for infinite sums of strictly proper scores is
expressed by the following result.
\begin{theorem}\label{thm:1}
Let ${\cal C}$ be a collection of strictly proper scoring rules of
the form~(\ref{eq:gdef}) that satisfies the uniform spread condition.
Let $P$ be a coherent$_3$ prevision defined over a collection $\D$ of
random variables that contains all of the random variables mentioned
in the statement of this theorem. Let $\{X_i\}_{i=1}^\infty$ be
random variables in $\D$ with
coherent$_3$ forecasts $P(X_i) = p_{i}$ for $i=1,2,\ldots.$ Assume
that the forecast for $X_i$ will be scored by a scoring rule
$g_i\in{\cal C}$ for each $i$. Finally, assume that
\begin{eqnarray}
\label{eq9} P \Biggl[\sum_{i=1}^\infty\mid
X_{i}-p_i \mid \Biggr] &=& V < \infty\quad\mbox{and}
\\
\label{eq2} P \Biggl[\sum_{i=1}^\infty
g_i(X_i,p_i) \Biggr] &= & W < \infty.
\end{eqnarray}
There does not exist a rival set of forecasts $\{q_i\}_{i=1}^\infty$
such that, for all $\omega\in\Omega$,
\begin{equation}
\label{eq:dom} \sum_{i=1}^\infty
g_i\bigl(X_i(\omega),p_i\bigr)>\sum
_{i=1}^\infty g_i
\bigl(X_i(\omega),q_i\bigr).
\end{equation}
\end{theorem}
Theorem~\ref{thm:1} asserts conditions under which infinite sums of
strictly proper scores, with coherent$_3$ forecasts $\{p_{i}\}
_{i=1}^\infty$ for
$\{X_{i}\}_{i=1}^\infty$, have no rival forecasts that
simply dominate, let alone uniformly strictly
dominate $\{p_i\}_{i=1}^\infty$. That is, even countably many
unconditional coherent$_3$ forecasts cannot be simply dominated
under the conditions of Theorem~\ref{thm:1}.
\begin{example}[(Example~\ref{examp:1} continued)]\label{examp:3}
Recall that
$\Omega= \{\omega_{i}\dvtx i = 1, \ldots\}$ is a countable space.
Consider the special case in which $P$ is a purely finitely
additive probability satisfying
$P(\{\omega_{i}\}) = p_{i} = 0$, for all $i$. So, $\sum_{i=1}^\infty
p_{i} = 0 < 1$, and $c=0$ in the notation of
Example~\ref{examp:1}. As before, let $W_{i}$ ($i = 1, \ldots$) be
the indicator functions for the states in $\Omega$. So $P(W_i)
= p_{i} = 0$ and combining the losses $W_{i} - p_{i} =W_{i}$, for
$i = 1, \ldots$ results in a uniform
sure-loss of $1$. But this example, with each $g_i$ equal to Brier
score times $\alpha_i>0$, satisfies
the conditions of Theorem~\ref{thm:1}, if the $\alpha_i$ are
bounded above. That is, there are no
rival forecasts $\{q_i\}_{i=1}^\infty$ for the
$\{W_{i}\}_{i=1}^\infty$ that simply dominate the
forecasts $\{p_i\}_{i=1}^\infty$ by weighted sum of Brier scores, let alone
uniformly strictly dominating these forecasts.
We can illustrate the conclusion of Theorem~\ref{thm:1} directly in
this example. The weighted sum of Brier
scores for the $p_i$ forecasts is\looseness=-1
\[
S(\omega)=\sum_{i=1}^\infty
\alpha_iW_i(\omega)^2\leq\sup
_i\alpha_i.
\]\looseness=0
Let $\{q_i\}_{i=1}^\infty$ be a rival set of
forecasts with $q_i\ne p_i$ for at least one $i$. The corresponding
weighted sum of Brier scores is
\begin{equation}
\label{eq:exampsum} \sum_{i=1}^\infty
\alpha_i\bigl[W_i(\omega)-q_i
\bigr]^2=S(\omega) -2\sum_{i=1}^\infty
\alpha_iq_iW_i(\omega)+\sum
_{i=1}^\infty\alpha_iq_i^2.
\end{equation}
Let $d=\sum_{i=1}^\infty\alpha_iq_i^2$, which must be strictly greater
than 0.
Define $i(\omega)$ to be the unique value of $i$ such that $W_i(\omega)=1$.
The right-hand side of (\ref{eq:exampsum}) can then be written as
$S(\omega)-2\alpha_{i(\omega)}q_{i(\omega)}+d$. If $d=\infty$, then
the rival forecasts clearly fail to dominate the original forecasts.
If $d<\infty$, then $\lim_{i\rightarrow\infty}\sqrt{\alpha_i}q_i=0$.
Because the $\alpha_i$ themselves are bounded, it follows that all but finitely
many $\alpha_i|q_i|$ are less than $d/2$. For each $\omega$ such that
$\alpha_{i(\omega)}|q_{i(\omega)}|<d/2$, we have the weighted sum of
Brier scores displayed in (\ref{eq:exampsum})
strictly greater than $S(\omega)$, hence the rival forecasts do not
dominate the original forecasts.
\end{example}
Theorem~\ref{thm:1}, as illustrated by Example~\ref{examp:3}, shows
that the modified decision problem in de Finetti's prevision game---modified
to include infinite sums of betting outcomes---is not
isomorphic to the modified forecasting problem under strictly proper
scoring rules---modified to include infinite sums of scores. In particular,
abstaining from betting, which is the alternative that uniformly
dominates the losses for coherence$_1$, is not an available alternative
under forecasting with strictly proper scores. In summary, the two
criteria, coherence$_1$ and coherence$_3$ behave differently when
probability is merely finitely additive and we try to combine
countably many forecasts.\looseness=1
We conclude this section with an example to show why we assume that
the class of scoring rules satisfies the uniform spread condition in
Theorem~\ref{thm:1}.
\begin{example}
This example satisfies all of the conditions of Theorem~\ref{thm:1}
except that the class of scoring rules fails
the uniform spread condition. We show that the conclusion to
Theorem~\ref{thm:1} also fails. For each integer $i\geq1$, let
$\alpha_i=2^{-i-1}$, and define
\[
f_i(v)=\cases{
2,&\quad $\mbox{if $v\leq
\alpha_i$,}$
\vspace*{2pt}\cr
\displaystyle\frac{2}{\alpha_i}, &\quad
$\mbox{if $v>\alpha_i$.}$}
\]
Let $\lambda_i$ be the measure whose Radon--Nikodym derivative with
respect to Lebesgue measure is $f_i$, and define $g_i$ by
(\ref{eq:gdef}) using $\lambda=\lambda_i$. The form of $g_i$ is as follows:
\begin{eqnarray*}
g_i(x,q)=\cases{
(x-q)^2,&\quad
$\mbox{if $x, q\leq\alpha_i$,}$
\vspace*{2pt}\cr
\displaystyle(x-\alpha_i)^2+\frac{1}{\alpha_i} (q-
\alpha_i)^2+\frac{2}{\alpha
_i}(q-\alpha _i) (
\alpha_i-x), &\quad $\mbox{if $x\leq\alpha_i\leq q$,}$
\vspace*{2pt}\cr
\displaystyle\frac{1}{\alpha_i}(x-\alpha_i)^2+(q-
\alpha_i)^2+2(\alpha _i-q) (x-
\alpha_i) ,&\quad$\mbox{if $q\leq\alpha_i\leq x$,}$
\vspace*{2pt}\cr
\displaystyle\frac{1}{\alpha_i}(x-q)^2,&\quad $\mbox{if $\alpha_i\leq x, q$.}$}
\end{eqnarray*}
These scoring rules fail
the uniform spread condition because arbitrarily short intervals with
both endpoints positive have arbitrarily large $\lambda_i$ measure as
$i$ increases. Let $\{A_i\}_{i=1}^\infty$ be a partition of the real
line, and let $P(\cdot)$ be a finitely additive probability such that
$P(A_i)=0$ for all $i$. For each integer $i\geq1$, let $p_i=2^{-i}$ and
$q_i=2^{-i-1}$, and define
\[
X_i(\omega)=\cases{
p_i,&\quad
$\mbox{if $\omega\in A_i^C$,}$
\vspace*{2pt}\cr
q_i-1,&\quad $\mbox{if $\omega\in A_i$.}$}
\]
It follows that $P(X_i)=p_i$ for all $i$, and
\[
P \Biggl[\sum_{i=1}^\infty|X_i-p_i|
\Biggr]=P \Biggl[ \sum_{i=1}^\infty
A_i|q_i-1-p_i| \Biggr]=1,
\]
so that (\ref{eq9}) holds. Next, compute the various scores:
\begin{eqnarray*}
&&g_i\bigl(X_i(\omega),p_i
\bigr)\\
&&\qquad=A_i(\omega) \biggl\{ (q_i-1-
\alpha_i)^2+\frac{1}{\alpha_i}(p_i-
\alpha_i)^2+ \frac{2}{\alpha_i}(p_i-
\alpha_i) (\alpha_i-q_i+1) \biggr\}
\\
&&\qquad=A_i(\omega)\bigl[3+2^{-i-1}\bigr],
\\
&&g_i\bigl(X_i(\omega),q_i
\bigr)\\
&&\qquad=A_i^C(\omega)\frac{1}{\alpha_i}(p_i-q_i)^2
+A_i(\omega) (q_i-1-q_i)^2
\\
&&\qquad=A_i^C(\omega)2^{-i-1}+A_i(
\omega).
\end{eqnarray*}
Define $i(\omega)=i$ for that unique $i$ such that $\omega\in A_i$.
When we sum up the scores for the forecasts $\{p_i\}_{i=1}^\infty$, we get
\[
\sum_{i=1}^\infty g_i
\bigl(X_i(\omega),p_i\bigr)=\sum
_{i=1}^\infty A_i(\omega)
\bigl[3+2^{-i-1}\bigr]=3+2^{-i(\omega)-1}.
\]
It follows that $P(\sum_{i=1}^\infty g_i(X_i,p_i))=3$, so (\ref{eq2}) holds.
The sum of the $\{q_i\}_{i=1}^\infty$ scores is
\[
\sum_{i=1}^\infty g_i
\bigl(X_i(\omega),q_i\bigr)=\sum
_{i=1}^\infty \bigl[A_i^C(
\omega)2^{-i-1} +A_i(\omega)\bigr]=1.5-2^{-i(\omega)-1}.
\]
Finally, compute the difference in total scores:
\[
\sum_{i=1}^\infty g_i
\bigl(X_i(\omega),p_i\bigr)-\sum
_{i=1}^\infty g_i\bigl(X_i(
\omega),q_i\bigr) =1.5+2^{-i(\omega)}>1.5,
\]
hence the scores of the $\{p_i\}_{i=1}^\infty$ forecasts are uniformly
strictly dominated by the scores of a set of rival forecasts.
\end{example}
\subsection{Dominance for countable sums of conditional
forecasts}\label{sub:dominance4}
Definition~\ref{def:condcoh} allows mixing conditional forecasts with
unconditional forecasts by setting $P(X|H)=P(X)$ whenever $H=\Omega$.
De Finetti showed that, if $P(X)$, $P(X|H)$ and $P(HX)$ are all
specified, a necessary condition for coherence$_1$ is
that
\begin{equation}
\label{eq:cond} P(HX)=P(H)P(X|H),
\end{equation}
so that $P(X|H)$ is the usual conditional expected value of $X$ given
$H$ whenever $P(H)>0$. For this reason, conditional forecasts are
often called \emph{conditional expectations}.
The concept of \textit{conglomerability} plays a central role in our
results about coherence for combining countably many conditional
forecasts. Conglomerability in a partition $\pi=
\{H_j\dvtx j\in J \}$ of conditional expectations $P(\cdot
\mid H_j)$ over a class $\D$ of random variables $X$ is the
requirement that the unconditional expectation of each $X\in\D$
lies within the range of its conditional expectations given
elements of $\pi$.
\begin{definition}\label{def:nonc} Let $P$ be a finitely additive
prevision on a set $\D$ of random variables,
and let $\pi=\{H_j\dvtx j\in J\}$ be a partition of $\Omega$ such that
conditional prevision $P(\cdot|H_j)$ has been defined for all
$j$. If, for each $X\in\D$,
\[
\inf_{j \in J} P(X \mid H_{j}) \leq P(X) \leq\sup
_{j \in J} P(X \mid H_{j}),
\]
then $P$ is
\emph{conglomerable in the partition $\pi$ with respect to $\D$}.
Otherwise, $P$ is \emph{nonconglomerable} in $\pi$ with respect to
$\D$.
\end{definition}
If a decision maker's coherent$_1$ or coherent$_3$
forecasts fail conglomerability in a partition $\pi$,
Theorem~\ref{thm:2} below shows
there exist countably many conditional forecasts that are uniformly
strictly dominated.
On the other hand, if the decision maker's previsions for random
variables satisfy a condition (see Definition~\ref{def:ltp}) similar
to being conglomerable in $\pi$, Theorem~\ref{thm:3} below
establishes that no countable set of forecasts,
conditional on elements of $\pi$, can be uniformly strictly dominated.
What we mean by ``similar'' is explained in Section~\ref{sec:cdltp} below.
\begin{theorem}\label{thm:2}
Let $P$ be a finitely additive prevision, and let $\D$ be a set of
random variables. Let $\pi= \{H_j\}_{j=1}^\infty$ be a denumerable
partition and let $P(\cdot\mid H_j)$ be the corresponding conditional
previsions associated with $P$. Let ${\cal C}$ be a collection of strictly
proper scoring rules of the form (\ref{eq:gdef}) that satisfies the
uniform similarity condition.
Assume that the conditional previsions $P(\cdot\mid H_j)$
are nonconglomerable in $\pi$ with respect to $\D$. Then there exists
a random variable $X\in\D$ with $p_X=P(X)$ and $p_j=P(X|H_j)$ for all
$j$ such that
{(2.1)} the countable sum
\[
\alpha_0(X-p_X)+\sum_{j=1}^\infty
\alpha_jH_j(X-p_j),
\]
of individually fair options is uniformly strictly dominated by
abstaining, and
{(2.2)}
if the forecast for $X$ is scored by $g_0\in{\cal C}$ and the
conditional forecast for $X$ given $H_j$ is scored by $g_j\in{\cal C}$
for $j=1,2,\ldots,$ then the sum of the scores,
\[
g_0\bigl(X(\omega),p_X\bigr)+\sum
_{j=1}^\infty H_j(\omega)g_j
\bigl(X(\omega),p_j\bigr),
\]
is uniformly strictly dominated by the sum of scores from a rival set
of forecasts.
\end{theorem}
We illustrate Theorem~\ref{thm:2} with an example of
nonconglomerability due to \mbox{\citet{dubins1975}}. This example is
illuminating as the conditional probabilities do not involve
conditioning on null events.
\begin{example}\label{examp:2}
Let $\Omega= \{\omega_{ij}\dvtx i = 1,2; j = 1, \ldots\}$. Let $F =
\{\omega_{2j}, j = 1, \ldots\}$ and let $H_{j} = \{\omega_{1j},
\omega_{2j}\}$. Define a merely finitely additive probability $P$ so
that $P(\{\omega_{1j}\}) = 0, P(\{\omega_{2j}\}) = 2^{-(j+1)}$ for $j
= 1, \ldots,$ and let $P(F) = p_{F} = 1/2$. Note that $P(H_{j}) =
2^{-(j+1)} > 0$, so $P(F \mid
H_{j}) = 1 = p_{j}$ is well defined by the multiplication rule for
conditional probability. Evidently, the conditional probabilities
$\{P(F \mid H_{j})\}_{j=1}^\infty$ are nonconglomerable in $\pi$
since $P(F) = 1/2$ whereas $P(F \mid H_{j}) = 1$ for all $j$.
For {(2.1)}, Consider the fair options $\alpha_{j}H_{j}(F -
p_{j})$ for $j = 1, \ldots$ and $\alpha_F(F - p_F)$.
Choose $\alpha_j = 1$ and $\alpha_F=-1$. Then
\begin{eqnarray*}
&&\Biggl[-\bigl(F(\omega) - p_F\bigr) + \sum
_{j=1}^\infty H_j(\omega)\bigl[F(\omega) -
p_j\bigr] \Biggr] \\
&&\qquad= \cases{
0.5 - 1.0 = -0.5,&\quad $\mbox{if $\omega\notin F$,}$
\vspace*{2pt}\cr
-0.5 + 0.0 = -0.5,&\quad $\mbox{if $\omega\in F$.}$}
\end{eqnarray*}
Hence, these infinitely many individually fair options
are not collectively fair when taken together. Their sum is
uniformly strictly dominated by $0$ in $\Omega$, corresponding to the
option to abstain from betting.
Regarding {(2.2)}, unlike the situation with
Theorem~\ref{thm:1} involving countably many unconditional forecasts,
the sum of Brier scores from these conditional forecasts are
uniformly strictly dominated. In particular, the sum of Brier scores
for these forecasts is
\begin{eqnarray*}
&&\bigl(F(\omega) - p_F\bigr)^2 + \sum
_{j=1}^\infty H_j(\omega)\bigl[F(
\omega)-p_j\bigr]^2\\
&&\qquad= \cases{
0.25 + 1.00 = 1.25,&\quad $\mbox{if $\omega\notin F$,}$
\vspace*{2pt}\cr
0.25 + 0.00 = 0.25,&\quad $\mbox{if $\omega\in F$.}$}
\end{eqnarray*}
Consider the rival forecasts $Q(F \mid H_{j}) = 0.75 = q_{j}$ and
$Q(F) = 0.75 = q_F$. These correspond to the
countably additive probability $Q(\{\omega_{1j}\}) = 0.25 \times
2^{-j}$ and $Q(\{\omega_{2j}\}) = 0.75 \times2^{-j}$ for $j = 1,
\ldots.$ Then the combined Brier score from these countably many
rival forecasts is
\begin{eqnarray*}
&&\bigl(F(\omega)-q_F\bigr)^2+\sum
_{j=1}^\infty H_j(\omega)\bigl[F(
\omega)-q_j\bigr]^2 \\
&&\qquad=\cases{
9/16 + 9/16 = 1.125,&\quad $\mbox{if $\omega\notin F$,}$
\vspace*{2pt}\cr
1/16 + 1/16 = 0.125,&\quad $\mbox{if $\omega\in F$,}$}
\end{eqnarray*}
which is 0.125 less than the sum of the Brier scores of the original forecasts.
\end{example}
We offer one more example to show why we assume that the class of
scoring rules satisfies the uniform similarity condition in
Theorem~\ref{thm:2}.\vspace*{-1pt}
\begin{example}[(Example~\ref{examp:2} continued)]
Recall that we have a partition $\pi=\{H_j\}_{j=1}^\infty$ and an
event $F$ with $p_F=P(F)=0.5$ and $p_j=P(F|H_j)=1$ for all~$j$.
Let the unconditional forecast for $F$ be scored by Brier score, and let
the conditional forecast for $F$ given $H_j$ be scored by
$2^{-j-1}$ times Brier score. These scoring rules fail the
uniform similarity condition. We show that the conclusion to
Theorem~\ref{thm:2} fails. Specifically, we show that there is no
rival set of forecasts $q_F$ for $F$ and $q_j$ for $H_j$
($j=1,2,\ldots$) whose sum of scores uniformly strictly dominates the
original forecasts.
The total of the scores for the original forecasts is
\begin{equation}
\label{eq:pscore} \frac{1}{4}+\sum_{j=1}^\infty
H_j(\omega)2^{-j-1}\bigl[1-F(\omega)\bigr]^2.
\end{equation}
Consider an arbitrary rival set of forecasts with $q_F$ for $F$ and
$q_j$ for $F$ conditional on $H_j$. The sum of the scores for the rival
forecasts is
\begin{equation}
\label{eq:qscore} \bigl[q_F-F(\omega)\bigr]^2+\sum
_{j=1}^\infty H_j(\omega)2^{-j-1}
\bigl[q_j-F(\omega)\bigr]^2.
\end{equation}
Let $i(\omega)=j$ when $\omega\in H_j$. Then the difference
(\ref{eq:pscore}) minus (\ref{eq:qscore}) is
\begin{equation}
\label{eq:scorediff} \tfrac{1}{4}-\bigl[q_F-F(\omega)
\bigr]^2 +2^{-i(\omega)-1} \bigl(\bigl[1-F(\omega)
\bigr]^2-\bigl[q_j-F(\omega)\bigr]^2 \bigr).
\end{equation}
If $q_F=0.5$, then (\ref{eq:scorediff}) becomes
\begin{equation}
\label{eq:scorediff2} 2^{-i(\omega)}(1-q_{i(\omega)}) \biggl[\frac{1+q_{i(\omega
)}}{2}-F(
\omega ) \biggr].
\end{equation}
If there exists $\omega$ such that $q_{i(\omega)}\geq1$, (\ref
{eq:scorediff2})
is nonpositive, and the rival forecasts do not strictly dominate. If
all $q_{i(\omega)}<1$, (\ref{eq:scorediff2}) is negative for all
$\omega\in F$, and there is no dominance. If $q_F\ne0.5$, then
(\ref{eq:scorediff}) is at most
\begin{equation}
\label{eq:scorediff3} \tfrac{1}{4}-\bigl[q_F-F(\omega)
\bigr]^2+2^{-i(\omega)}.
\end{equation}
No matter what $q_F\ne0.5$ we pick, either $(q_F-1)^2$ or $(q_F-0)^2$ is
greater than $1/4$. Let $\delta=1/4-\max\{[q_F-1]^2,q_F^2\}$. For
$j>-\log_2(\delta)$, (\ref{eq:scorediff3}) is negative either for all
$\omega\in F\cap H_j$ or all $\omega\in F^C\cap H_j$. So, there is no
dominance.\vspace*{-1pt}
\end{example}
Last, we establish conditions under which combining strictly proper
scores from countably many conditional forecasts given elements of a
partition, or combining
the losses from countably many fair options based on those forecasts,
does not result in a uniform sure loss. A definition is useful first.
\begin{definition}\label{def:ltp}
Let $P$ be a finitely additive prevision on a set $\D$ of
random variables,
and let $\pi=\{H_j\dvtx j\in J\}$ be a\vadjust{\goodbreak} partition of $\Omega$ such that
conditional prevision $P(\cdot|H_j)$ has been defined for all
$j$. For each random variable $X\in\D$, we let $P(X|\pi)$ denote
the random variable $Y$ defined by $Y(\omega)=P(X|H_j)$ for all
$\omega\in H_j$ and all $j$. We say that $P$ \emph{satisfies the law
of total previsions in $\pi$ with respect to $\D$} provided that
for each random variable $X\in\D$, $P(X) = P[P(X \mid\pi)]$.
\end{definition}
\begin{theorem}\label{thm:3}
Let $P$ be a finitely additive prevision, and let $\D$ be a set of
random variables such that $P$ satisfies the law of
total previsions in $\pi= \{H_j\}_{j=1}^\infty$ with respect to $\D$.
Let $X\in\D$ be a random variable with
finite prevision $p_X=P(X)$ and finite conditional prevision
$p_j=P(X|H_j)$ given each $H_j$. Assume that $P(\cdot|H_j)$ is a finitely
additive expectation (in the sense of Definition~\ref{def:daniell})
that satisfies $P(X|H_j)=P(H_jX|H_j)$ for every $j$. Let ${\cal C}$ be
a collection of
strictly proper scoring rules.
{(3.1)}
Let $\{\alpha_j\}_{j=0}^\infty$ be real numbers. The sum of losses
\begin{equation}
\label{eq18} \alpha_0\bigl(X(\omega) - p_X\bigr) +
\sum_{j=1}^\infty\alpha_jH_j(
\omega)\bigl[X(\omega) - p_j\bigr],
\end{equation}
is not uniformly strictly dominated by abstaining.
{(3.2)} Let $g_0,g_1,\ldots$ be elements of ${\cal C}$. There is
no rival set of forecasts that uniformly strictly dominates the sum of scores
\begin{equation}
\label{eq24} g_0\bigl(X(\omega),p_X\bigr)+\sum
_{j=1}^\infty H_j(
\omega)g_j\bigl(X(\omega),p_j\bigr).
\end{equation}
\end{theorem}
\subsection{Conglomerability, disintegrability and the law of total
previsions}\label{sec:cdltp}
We claimed earlier that the law of total previsions in a partition
$\pi$ is similar to conglomerability in $\pi$. The claim begins with
a result of \citet{dubins1975}.
Dubins defines \emph{conglomerability in partition} $\pi$ of a finitely
additive prevision $P$ by the requirement that, for all bounded random
variables $X$,
\[
\mbox{if } \forall H \in\pi P(X |H) \geq0, \mbox{then } P(X) \geq0.
\]
Dubins' definition of conglomerability in $\pi$ is equivalent to
Definition~\ref{def:nonc} with respect to the set of all bounded
random variables. However, for a set $\D$ that includes unbounded random
variables and/or does not include all bounded random variables,
the two definitions are not equivalent without further assumptions.
Definition~\ref{def:nonc} is based on the definition given by de Finetti
[(\citeyear{deFinetti1974}), Section~4.7], which generalizes to unbounded random variables
more easily.
\citet{dubins1975} also defines \emph{disintegrability of $P$ in partition}
$\pi$
by the requirement that, for every bounded random variable $X$,
\[
P(X) = \int P(X |h) \,dP(h),
\]
where the finitely additive integral is as developed by Dunford and
Schwartz [(\citeyear{dunford-schwartz1958}), Chapter III].
Moreover, he establishes that conglomerability and disintegrability
in $\pi$ are equivalent for the class of bounded random
variables.\looseness=1
The law of total previsions in Definition~\ref{def:ltp}, with respect
to the set of all bounded random variables, is equivalent to
disintegrability in Dubins' sense, but not necessarily for sets that either
include some unbounded random variables or fail to include some bounded
random variables. In addition, not all real-valued coherent$_1$
previsions admit an integral
representation in the sense of Dunford and Schwartz for sets that
include unbounded random variables. For discussion of
the problem and related issues, see \citet{berti-etal2001};
Berti and Rigo
(\citeyear{berti-rigo1992,berti-rigo2000,berti-rigo2002}); Schervish, Seidenfeld,
and Kadane (\citeyear{ssk2008a}) and
\citet{ssk2009a}. As described in
Appendix \ref{app:lf}, we use a definition of
finitely additive integral that is a natural extension of coherent$_1$
prevision. In this way, the law of total previsions
extends Dubins' definition of disintegrability from bounded to
unbounded random variables without introducing the technical details
of Dunford and Schwartz. Finally, Theorem~1 of \citet{ssk2008a} gives conditions under which
conglomerability (Definition~\ref{def:nonc}) is equivalent to the law
of total previsions. The following is a translation of that result into
the notation and terminology of the present paper.
\begin{theorem}\label{thm:2008b}
Let $P$ be a finitely additive prevision on a set $\D$ of
random variables. Let $\pi= \{H_j\}_{j=1}^\infty$ be a denumerable
partition and let $P(\cdot\mid H_j)$ be the corresponding conditional
previsions associated with $P$. Assume that, for all $j$,
$P(\cdot|H_j)$ is a finitely additive expectation on $\D$. Also assume
that, for all $X\in\D$:
\begin{itemize}
\item$P(X)$ is finite,
\item$P(X|H_j)$ is finite for all $j$,
\item$H_jX\in\D$ for all $j$,
\item$P(H_jX|H_j)=P(X|H_j)$ for all $j$, and
\item$X-Y\in\D$, where $Y$ is defined (in terms of $X$) in
Definition~\ref{def:ltp}.
\end{itemize}
Then $P$ is conglomerable in $\pi$ with
respect to $\D$ if and only if $P$ satisfies the law of total
previsions in $\pi$ with respect to $\D$.
\end{theorem}
Under the conditions of Theorem~\ref{thm:2008b},
Theorems \ref{thm:2} and \ref{thm:3} show that, when the conditioning
events form a countable partition $\pi$, coherence$_1$ and
coherence$_3$ behave the same when extended to include, respectively,
the countable sum of individually fair options, and
the total of strictly proper scores from the forecasts. If and
only if these
coherent quantities are based on conditional expectations that are
conglomerable in $\pi$, then no failures of the Dominance Principle
result by combining infinitely many of them.
\citet{SchervishSeidenfeldandKadane1984} show that each merely
finitely additive probability fails to be conglomerable in some
countable partition. But each countably additive probability
has expectations that are conglomerable in each countable partition.
Thus, the conjunction of Theorems \ref{thm:1},
\ref{thm:2} and \ref{thm:3} identifies where the debate whether personal
probability may be merely finitely additive runs up against the debate
whether to extend either coherence criterion in order to apply it with
countable combinations of quantities.
We arrive at the following conclusions:
\begin{itemize}
\item Unless unconditional coherent$_{1}$ forecasts arise from a
countably additive probability,
combining countably many unconditional coherent$_{1}$ forecasts
into a single option may be dominated by abstaining.
\item However, under the conditions of Theorem~\ref{thm:1}, strictly
proper scoring rules are not similarly affected. The scores from
countably many coherent$_3$ unconditional
forecasts may be summed together without leading to a violation of
the Dominance Principle.
\item Unless conditional forecasts arise from
a set of conglomerable conditional probabilities, the
Dominance Principle does not allow combining countably many of these
quantities into a single option. Hence, only countably additive
conditional probabilities satisfy the Dominance Principle when an
arbitrary countable set of conditional quantities are summed
together.
\end{itemize}
\section{Incentive compatible elicitation of infinitely many forecasts
using strict\-ly proper scoring rules}\label{sec:proper}
Scoring an agent based on the values of the fair gambles constructed
from coherent$_1$ forecasts, is not proper. Because of the
presence of the opponent in the game, who gets to choose whether to
buy or to sell the random variable $X$ at the decision maker's
announced price, the decision maker faces a strategic choice of
pricing. For example, if the decision maker suspects that the
opponent's fair price, $Q(X)$, is greater than his own, $P(X)$, then
it pays to inflate the announced price and to offer the opponent,
for example, $R(X) = [P(X) + Q(X)]/2$, rather than offering $P(X)$.
Thus, the forecast-game as de Finetti defined it for coherence$_1$
is not incentive compatible for eliciting the decision maker's fair
prices.
With a finite set of forecasts and a strictly proper scoring rule
for each one, using the finite sum of the scores as the score
for the finite set preserves strict propriety. That is, with the sum
of strictly proper scores as the score for the finite set, a coherent forecaster
minimizes the expected sum of scores by minimizing each one, and this
solution is unique.
Here, we report what happens to the propriety of strictly proper scores
in each
of the three settings of the three theorems presented in
Section~\ref{sec:extensions}. That is, we answer the question whether
or not, in each of these three settings, the coherent forecaster
minimizes expected score for the infinite sum of strictly proper scores by
announcing her/his coherent forecast for each of the infinitely many
variables. These findings are corollaries to the respective
theorems.
\begin{corollary}
Under the assumptions of Theorem~\ref{thm:1}, the infinite sum of
scores applied to the infinite set of forecasts $\{p_i\}_{i=1}^\infty$
is a
strictly proper scoring rule.
\end{corollary}
\begin{corollary}
Under the assumptions used for {(2.2)} of Theorem~\ref{thm:2},
namely when
the conditional probabilities $P(F \mid H_j) = p_j$ are
nonconglomerable in $\pi$, then the infinite
sum of strictly proper scores applied to the infinite set of conditional
forecasts $\{p_j\}_{j=1}^\infty$ is not proper.
\end{corollary}
\begin{corollary}
Under the assumptions used to establish {(3.2)} of Theorem~\ref{thm:3},
namely that $P$ satisfies the law of total previsions in $\pi$, the
infinite sum of
strictly proper scores applied to the infinite set of conditional forecasts
$\{p_j\}_{j=1}^\infty$ is a proper scoring rule.
\end{corollary}
Thus, these results about the propriety of infinite sums of strictly proper
scores parallel the respective results about extending coherence$_3$
to allow infinite sums of scores.
\section{Summary}\label{sec:summary}
We study how two different coherence criteria
behave with respect to a Dominance Principle when countable
collections of random variables are included. Theorem~\ref{thm:1}
shows that, in
contrast with fair prices for coherence$_1$, when strictly proper
scores from infinitely many unconditional forecasts are summed together
there are no new failures of the Dominance Principle for
coherence$_3$. That is, if an infinite set of probabilistic forecasts
$\{p_{i}\}_{i=1}^\infty$ are even simply dominated by some rival
forecast scheme
$\{q_{i}\}_{i=1}^\infty$ in total score, then the $\{p_{i}\}
_{i=1}^\infty$ are not
coherent$_3$, that is, some finite subset of them is uniformly strictly
dominated in total score. However, because each merely finitely
additive probability fails to be conglomerable in some denumerable
partition, in the light of Theorem~\ref{thm:2}, neither of the two
coherence criteria discussed here may be relaxed in order to apply the
Dominance Principle with infinite combinations of conditional options.
Merely finitely additive probabilities then would become
\textit{incoherent}.
Specifically, the conjunction of
Theorems \ref{thm:1}--\ref{thm:2008b} shows that it matters only in
cases that involve
nonconglomerability whether incoherence$_3$ is established using
scores from a finite rather than from an infinite
combination of forecasts. In that one respect, we think
coherence$_3$ constitutes an improved version of the concept of
coherence. Coherence$_1$ applied to a merely finitely
additive probability leads to failures of the Dominance Principle
both with infinite combinations of unconditional and infinite
combinations of nonconglomerable conditional probabilities.
Coherence$_3$ leads to failures of the Dominance Principle only with
infinite combinations of nonconglomerable conditional
probabilities.\vadjust{\goodbreak}
A referee suggested that de Finetti might have been working with a
different Dominance Principle, here denoted Dominance*.
\begin{quote}
\emph{Dominance*}: Let $O_i$ and $O_j$ be two options in ${\cal O}$. If
$O_i$ uniformly (strictly) dominances $O_j$ and there exists an option
$O_k$ in ${\cal O}$ that is not itself dominated by some $O_t$ in
${\cal O}$, then $O_j$ is an inadmissible choice from ${\cal O}$.
\end{quote}
Dominance* requires that some option from ${\cal O}$ is
undominated if dominance signals inadmissibility. With respect to the
decision problems considered in this paper, each of our results
formulated with respect to the Dominance Principle obtains also with
Dominance*. Because Dominance* implies Dominance as we have
defined it, the only result that needs to be checked is
Theorem~\ref{thm:2}. In that case, so long as ${\cal O}$
contains options that correspond to a probability that satisfies the
law of total previsions in $\pi$ (as will all countably additive
probabilities) then Theorem~\ref{thm:3} says that such options will be
undominated. So, we could replace Dominance by Dominance* in the
results of this paper.
\begin{appendix}
\section{Finitely additive expectations}\label{app:lf}
This appendix gives the definitions of infinite prevision and finitely
additive expectation along with brief motivation for these definitions.
Details are given in
the supplemental article [Schervish, Seidenfeld and
Kadane (\citeyear{ssk2013})].
\subsection{Infinite previsions}\label{sec:infprev}
Our theorems assume that various random variables have finite
previsions. In the proof of Theorem~\ref{thm:1}, the possibility
arises that some other random variable has infinite prevision.
Definition~\ref{def:coh} makes no sense if infinite previsions are
possible. Fortunately, we can extend the concept of coherent$_1$
(conditional) prevision to handle infinite values, which correspond to
expressing a willingness either to buy or to sell a gamble, but not
both.
\begin{definition}\label{def:cohinf}
Let $\{P(X_i|B_i)\dvtx i\in I\}$ be a collection of conditional
previsions. The previsions are \emph{coherent$_1$} if, for every
finite $n$, every $\{i_1,\ldots,i_n\}\subseteq I$, all real
$\alpha_1,\ldots,\alpha_n$ such that $\alpha_j\leq0$ for all $j$
with $P(X_{i_j}|B_{i_j})=\infty$ and $\alpha_j\geq0$ for all $j$ with
$P(X_{i_j}|B_{i_j})=-\infty$, and all real $c_1,\ldots,c_n$ such that
$c_j=P(X_{i_j}|B_{i_j})$ for each $j$ such that $P(X_{i_j}|B_{i_j})$
is finite, we have
\begin{equation}
\label{eq:yacc} \inf_{\omega\in\Omega}\sum_{j=1}^n
\alpha_jB_{i_j}(\omega )\bigl[X_{i_j}(\omega
)-c_j\bigr]\leq0.
\end{equation}
That is, no linear combination of gambles may be uniformly strictly
dominated by the alternative option of abstaining.
\end{definition}
Notice the restrictions on the signs of coefficients in
Definition~\ref{def:cohinf}, namely that for each infinite prevision,
$\alpha_j$ has the opposite sign as the prevision. These restrictions
express the meaning of infinite
previsions as being \emph{one-sided} in the
sense that they merely specify that all real numbers are either
acceptable buy prices (for $\infty$ previsions) \emph{or} acceptable
sell prices
(for $-\infty$ previsions) but not fair prices for both transactions.
\citet{crisma97} and \citet{crisma01}
give alternate definitions of coherence for infinite previsions and
conditional previsions. But their definition does not make
clear the connection to gambling. However, the definition of Crisma, Gigante and
Millossovich (\citeyear{crisma97}) and Definition~\ref{def:cohinf} are equivalent for unconditional
previsions, as shown in the supplemental article [Schervish, Seidenfeld and
Kadane (\citeyear{ssk2013})].
\subsection{Prevision and expectation}\label{sec:prevexp}
Throughout this paper, an expectation with respect to a finitely
additive probability will be defined as a special type of linear
functional on a space of random variables. [See \citet{heathsudderth1978} for the case of bounded random variables.] Infinite
previsions are allowed in the sense of Section~\ref{sec:infprev}.
\begin{definition}\label{def:daniell}
Let $\L$ be a linear space of real-valued functions defined on
$\Omega$ that contains all constant functions, and
let $L$ be an extended-real-valued functional defined on $\L$.
If ($X,Y\in\L$ and $X\leq Y$) implies $L(X)\leq L(Y)$, we say that
$L$ is
\emph{nonnegative}.
We call $L$ an \emph{extended-linear functional} on $\L$, if, for all
real $\alpha,\beta$ and all $X,Y\in\L$,
\begin{equation}
\label{eq:el} L(\alpha X+\beta Y)=\alpha L(X)+\beta L(Y),
\end{equation}
whenever the arithmetic on the right-hand side of (\ref{eq:el}) is
well defined (i.e., not $\infty-\infty$) and where $0\times\pm
\infty=0$
in (\ref{eq:el}).
A nonnegative extended-linear functional is called a \emph{finitely
additive Daniell integral}. [See Schervish, Seidenfeld and
Kadane (\citeyear{ssk2008b}).]
If $L(1)=1$, we say that $L$ is \emph{normalized}. A~normalized
finitely additive Daniell integral is called a \emph{finitely
additive expectation}.
\end{definition}
Note that, if $\infty-\infty$ appears on the right-hand side of
(\ref{eq:el}), $L(\alpha X+\beta Y)$ still has a value, but the value
cannot be determined from (\ref{eq:el}).
Finitely additive expectations are essentially equivalent to
coherent$_1$ previsions, as we prove in the supplemental article.
Finitely additive expectations also behave like integrals in many
ways, as we explain in more detail in the supplemental article.
In particular, when the finitely additive expectation defined here is
restricted to bounded functions, it is the
same as the definition of integral developed by \citet{dunford-schwartz1958}, and it is the same as the integral used by \citet{dubins1975} in
his results about disintegrability. Hence,
Definition~\ref{def:daniell} is an extension of the definition of
integral from sets of bounded functions to arbitrary linear spaces of
functions.
\section{Proofs of results}\label{app:aux}
\subsection{Proof of Lemma \texorpdfstring{\protect\ref{lem:sp}}{1}}
Let $g$ be of the form (\ref{eq:gdef}). Let $P$ be such that $p=P(X)$
is finite, and let $q_0$ be such that $P[g(X,q_0)]$ is finite.
If $q\ne p$, then
\begin{equation}
\label{eq:pdiff} P\bigl[g(X,q)-g(X,p)\bigr]=\lambda\bigl((q,p)\bigr)\bigl[p-r(q,p,
\lambda)\bigr],
\end{equation}
according to (\ref{eq:diff}). Because Lebesgue measure is absolutely
continuous with respect to $\lambda$, neither $\lambda((q,p))$ nor
$p-r(q,p,\lambda)$ equals 0 and they have the same sign. It follows that
(\ref{eq:pdiff}) is
strictly positive. Since $p$ is finite, (\ref{eq:pdiff}) is finite
with $q=q_0$, so that $P[g(X,p)]$ is also finite and so $q=p$ provides
the unique minimum value of $P[g(X,q)]$.
\subsection{Proof of Lemma \texorpdfstring{\protect\ref{lem:ubf}}{2}}
Since $r(b,a,\lambda)=r(a,b,\lambda)$, it suffices to assume that
$a<b$. Let $\varepsilon>0$.
(i) If $\lambda_i((a,b))\geq\varepsilon$ and $b_0<b$ is such
that $\lambda_i((a,b_0))=\varepsilon$, then the probability obtained by
normalizing $\lambda_i$ on the interval $(a,b)$ stochastically dominates
the probability obtained by normalizing $\lambda_i$ on the interval
$(a,b_0)$. Hence, $r(a.b,\lambda_i)\geq r(a,b_0,\lambda_i)$. So, it
suffices to find a $\delta$ that implies
$r(a,b,\lambda_i)-a\geq\delta$ for all $i\in I$ and all $a<b$
such that $\lambda_i((a,b))=\varepsilon$. For the remainder of the
proof, let $a<b$ with $\lambda_i((a,b))=\varepsilon$, and let $Q$ be the
probability obtained by normalizing $\lambda_i$ on $(a,b)$.
Let $\lambda_0$ be $U$ times Lebesgue measure. Then
$\lambda_0((a,a+\varepsilon/U))=\varepsilon$, and
$r(a,a+\varepsilon/U,\lambda_0)=a+\varepsilon/(2U)$. Because $f_i\leq U$, it
follows that $Q$ stochastically dominates the probability obtained by
normalizing $\lambda_0$ on $(a,a+\varepsilon/U)$, hence $r(a,b,\lambda
_i)\geq
a+\varepsilon/(2U)$, and $r(a,b,\lambda_i)-a\geq\varepsilon/(2U)$. The
proof $b-r(a,b,\lambda_i)\geq\varepsilon/(2U)$ is similar, so
$\delta_\varepsilon$ can be taken equal to $\varepsilon/(2U)$.
(ii) Let $i\in I$, and assume that $\lambda_i((a,b))\geq\varepsilon$.
Since $f_j(v)>L_if_i(v)$ for all $v$, we have
$\lambda_j((a,b))\geq L_i\varepsilon$, so $\gamma_{i,\varepsilon}$ can be
taken to be
$L_i\varepsilon$.
\subsection{Proofs of Theorem \texorpdfstring{\protect\ref{thm:1}}{1} and
Corollary \texorpdfstring{\protect\ref{thm:1}}{1}}
Because a larger random variable has a larger prevision than a smaller
random variable, a necessary condition for (\ref{eq:dom}) is that
\begin{equation}
\label{eq:qsum} Z=P \Biggl[\sum_{i=1}^\infty
g_i(X_i,q_i) \Biggr]\leq P \Biggl[\sum
_{i=1}^\infty g_i(X_i,p_i)
\Biggr]<\infty.
\end{equation}
Hence, we will assume that $Z<\infty$
from now on. Also, it is necessary for (\ref{eq:dom}) that $q_i\ne
p_i$ for at least one $i$, so we will assume this also.
In light of (\ref{eq:diff}), we can write, for each finite $k>0$,
\begin{eqnarray*}
\infty>Z-W&=&P \Biggl[\sum_{i=1}^\infty
g_i(X_i,q_i)-\sum
_{i=1}^\infty g_i(X_i,p_i)
\Biggr]
\\
&=&\sum_{i=1}^k\lambda_i
\bigl((q_i,p_i)\bigr)[p_i-r_i]
+P \Biggl[\sum_{i=k+1}^\infty
g_i(X_i,q_i)-\sum
_{i=k+1}^\infty g_i(X_i,p_i)
\Biggr]
\\
&\geq&\sum_{i=1}^k\lambda_i
\bigl((q_i,p_i)\bigr)[p_i-r_i]-W,
\end{eqnarray*}
where the inequality follows because $g_i$ is nonnegative for each $i$
and where $r_i = r(q_i,p_i,\lambda_i)$ from \eqref{eq:ra2}.
Since $Z-W$ does not depend on $k$, it follows that
$\sum_{i=1}^\infty\lambda_i((q_i,p_i))[p_i-r_i]$ is finite.
Because of (\ref{eq2}) and (\ref{eq:qsum}), the two series
$\sum_{i=1}^\infty g_i(X_i(\omega),q_i)$ and\break $\sum_{i=1}^\infty
g_i(X_i(\omega), p_i)$ are simultaneously finite with probability 1.
Let $B$ be the event that at least one of the two series is
finite. On $B^C$, both series sum to $\infty$, hence
(\ref{eq:dom}) fails unless $B^C=\varnothing$. Hence, we can assume
that $B=\Omega$ for the rest of the proof. It now follows that, for
all $\omega$,
\begin{eqnarray}
\label{eq:pq} &&\sum_{i=1}^\infty
g_i\bigl(X_i(\omega),q_i\bigr)-\sum
_{i=1}^\infty g_i
\bigl(X_i(\omega),p_i\bigr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad=\sum
_{i=1}^\infty\bigl[g_i\bigl(X_i(
\omega),q_i\bigr)-g_i\bigl(X_i(
\omega),p_i\bigr)\bigr].
\end{eqnarray}
We complete the proof by showing that
\begin{equation}
\label{eq5} P \Biggl[\sum_{i=1}^\infty
g_i(X_{i},q_{i})-\sum
_{i=1}^\infty g_i(X_{i},p_{i})
\Biggr]>0.
\end{equation}
Because a nonpositive random variable has nonpositive forecast,
(\ref{eq5}) implies that~(\ref{eq:dom}) cannot hold for all $\omega$.
In light of (\ref{eq:pq}), it suffices to show that
\begin{equation}
\label{eq:xyb} P \Biggl(\sum_{i=1}^\infty
\bigl[g_i(X_{i},q_{i})- g_i(X_{i},p_{i})
\bigr] \Biggr)>0.
\end{equation}
For each $k$,
\begin{equation}
\label{eq7} P \Biggl(\sum_{i=1}^k
\bigl[g_i(X_{i},q_{i})-g_i(X_{i},p_{i})
\bigr] \Biggr) = \sum_{i=1}^{k}
\lambda_i\bigl((q_i,p_i)\bigr)
(p_i-r_i) \geq0.
\end{equation}
Next, in light of (\ref{eq:diff}) and (\ref{eq7}), write
\begin{eqnarray}
\label{eq:split} &&P \Biggl(\sum_{i=1}^\infty
\bigl[g_i(X_i,q_i)-g_i(X_i,p_i)
\bigr] \Biggr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad= \sum_{i=1}^{k} \lambda_i
\bigl((q_i,p_i)\bigr) (p_i-r_i)
+
\nonumber
P \Biggl[\sum_{i=k+1}^{\infty}
\lambda_i\bigl((q_i,p_i)\bigr)
(X_i-r_i) \Biggr].
\end{eqnarray}
Since the left-hand side of (\ref{eq:split}) does not depend on $k$
and the
first sum on the right side is nondecreasing in $k$, it follows that
the second sum on the right-hand side is nonincreasing in $k$, and hence,
has a limit.
Let $T=\sum_{i=1}^\infty\lambda_i((q_i,p_i))(p_i-r_i)$, which is finite
and strictly positive (because $q_i\ne p_i$ for at least one $i$).
Then, the right-hand side of (\ref{eq:split}) becomes
\begin{equation}
\label{eq:sumT} T+\lim_{k\rightarrow\infty} P \Biggl[\sum
_{i=k+1}^\infty\lambda_i\bigl((q_i,p_i)
\bigr) (X_i-p_i) \Biggr].
\end{equation}
The proof will be complete if we can show that the limit in
(\ref{eq:sumT}) is 0.
First, we show that $\lim_{i\rightarrow\infty}\lambda_i((q_i,p_i))=0$.
If $\limsup_{i\rightarrow\infty}|\lambda_i((q_i,p_i))|= \ell>0$, then
there must exist a subsequence
$\{i_j\}_{j=1}^\infty$ with $|\lambda_{i_j}((q_{i_j},p_{i_j}))|>\ell
/2$ for
all $j$. For such a subsequence, the uniform spread condition implies
that there is $\delta_{\ell/2}>0$ such that $|p_{i_j}-r_{i_j}|\geq
\delta_{\ell/2}$. This would make $T=\infty$, a~contradiction.
It now follows that
\begin{eqnarray*}
\Biggl\llvert P \Biggl[\sum_{i=k+1}^\infty
\lambda _i\bigl((q_i,p_i)\bigr)
(X_i-p_i) \Biggr]\Biggr\rrvert &\leq& \max
_{i\geq k+1}\bigl|\lambda_i\bigl((q_i,p_i)
\bigr)\bigr|P \Biggl(\sum_{i=1}^\infty
|X_i-p_i| \Biggr)
\\
&=&V\max_{i\geq k+1}\bigl|\lambda_i\bigl((q_i,p_i)
\bigr)\bigr|,
\end{eqnarray*}
which can be made arbitrarily small by increasing $k$, and
(\ref{eq:xyb}) follows.
Corollary~\ref{thm:1} is equivalent to equation (\ref{eq5}), which is
established in the proof of Theorem~\ref{thm:1}.
\subsection{Proofs of Theorem \texorpdfstring{\protect\ref{thm:2}}{2} and
Corollary \texorpdfstring{\protect\ref{thm:2}}{2}}
Let $\pi= \{H_j\}_{j = 1}^\infty$ be a denumerable
partition. Nononglomerability means that there exists a random variable
$X$ such that either
\begin{eqnarray*}
\inf_j P(X|H_j) -P(X) &>& 0\quad\mbox{or}
\\
\sup_jP(X|H_j)-P(X)&<&0.
\end{eqnarray*}
Clearly, if $X$ satisfies one of
the above inequalities, $-X$ satisfies the other, hence we will
assume that the first inequality holds. Specifically, let
$p_X=P(X)$ and $p_j=P(X|H_j)$ for all $j$, and assume that
\[
\varepsilon= \inf_jp_j -p_X > 0.
\]
Also, for each $\omega\in\Omega$, let $i(\omega)$ be the unique
integer such
that $\omega\in H_{i(\omega)}$. Hence $H_j(\omega)=1$ if and only if
$j=i(\omega)$.
(2.1) Consider the following sum of individually fair options:
$X(\omega) - p_X$ and the countably many options
$-H_j(\omega)[X(\omega) - p_j]$ for $j=1,2,\ldots.$ Then, for each~$\omega$,
\begin{eqnarray*}
&&X(\omega) - p_X + \sum_{j=1}^\infty
-H_j(\omega)\bigl[X(\omega) - p_j\bigr]
\\
&&\qquad=X(\omega) - p_X - X(\omega) + p_{i(\omega)} =
-p_{X}+ p_{i(\omega)} \geq\varepsilon.
\end{eqnarray*}
Thus, the countable sum of the conditional forecasts for $X$ given
$H_j$, combined with the forecast for $X$ results in a
loss that is uniformly strictly dominated by 0.
(2.2) For an arbitrary set of forecasts $s_X$ for $X$ and $s_j$
for $X$ given
$H_j$ (for $j=1,\ldots$), the sum of the scores in state $\omega$ equals
\begin{eqnarray}
\label{eq16}
&&g_0\bigl(X(\omega),s_X
\bigr) + \sum_{j=1}^\infty H_j(
\omega )g_j\bigl(X(\omega ),s_j\bigr)\nonumber
\\
&&\qquad= g_0\bigl(X(\omega),s_X\bigr) +g_{i(\omega)}
\bigl(X(\omega), s_{i(\omega)}\bigr)
\\
&&\qquad=\int_{s_X}^{X(\omega)}\bigl[X(\omega)-v\bigr]\,d
\lambda_0(v) +\int_{s_{i(\omega)}}^{X(\omega)}\bigl[X(
\omega)-v\bigr]\,d\lambda_{i(\omega)}(v).
\nonumber
\end{eqnarray}
We can substitute the original forecasts $s_X=p_X$ and $s_j=p_j$,
$j=1,\ldots$
into~(\ref{eq16}) to obtain the total score
for each $\omega\in\Omega$. We can also identify dominating rival
forecasts $q_X$ and $q_j$, $j=1,\ldots,$ so that (\ref{eq16}) is
uniformly larger, for each state $\omega\in\Omega$ with $s_X=p_X$ and
$s_j=p_j$ than with $s_X=q_X$ and $s_j=q_j$.
Let $w_0=\lambda_0((p_X,p_X+\varepsilon))/2$, and let
$w_1=\gamma_{w_0}$, where $\gamma_{w_0}$ is from part (ii) of
Lemma~\ref{lem:ubf}. Let $q'$ be such that
$\lambda_0((q',p_X+\varepsilon))=w_0$. This makes
$\lambda_j((q',p_X+\varepsilon))\geq w_1$ for all $j$. For
each $j$, $p_j\geq p_X+\varepsilon$, so that $\lambda_j((q',p_j))\geq
w_1$. Let
$w_2=0.9\min\{w_0,w_1\}$, and let $q_j$ be such that
$\lambda_j((q_j,p_j))=w_2$ for all $j$. This makes $q_j> q'$ for all
$j$. Let $q_X$ be such that $\lambda_0((p_X,q_X))=w_2$. This makes $q_X<q'$.
We now form the difference between the scores for the original forecasts
and the rival forecasts. Subtracting (\ref{eq16}) with $s=q_X$ and
$s_j=q_j$ (for all $j$) from (\ref{eq16}) with $s=p_X$ and $s_j=p_j$
(for all $j$) yields
\begin{equation}
\label{eq17} \int_{p_X}^{q_X}\bigl[X(\omega)-v
\bigr]\,d\lambda_0(v)-\int_{q_{i(\omega
)}}^{p_{i(\omega)}}
\bigl[X(\omega)-v\bigr]\,d\lambda_{i(\omega)}(v).
\end{equation}
We need to find a positive number $\delta$ such that
(\ref{eq17}) is strictly greater than $\delta$ for all~$\omega$.
The difference in (\ref{eq17}) is greater than
\begin{eqnarray*}
&&\bigl[X(\omega)-q_X\bigr]\lambda_0
\bigl((p_X,q_X)\bigr)-\bigl[X(\omega)-q_{i(\omega)}
\bigr] \lambda_{i(\omega)}\bigl((q_{i(\omega)},p_{i(\omega)})\bigr)
\\
&&\qquad =w_2(q_{i(\omega)}-q_X)>w_2
\bigl(q'-q_X\bigr)>0.
\end{eqnarray*}
So, we set $\delta=w_2(q'-q_X)>0$, which
completes the proof.\vadjust{\goodbreak}
Corollary~\ref{thm:2} is immediate from {(2.2)} of
Theorem~\ref{thm:2}, as the existence of the rival set of dominating
forecasts, $\{q_j\}_{j=1}^\infty$, establishes that the forecaster
does not
minimize the infinite sum of expected scores by giving the forecast
$p_X$ and the conditional forecasts $\{p_j\}_{j=1}^\infty$.
\subsection{Proofs of Theorem \texorpdfstring{\protect\ref{thm:3}}{3} and
Corollary \texorpdfstring{\protect\ref{thm:3}}{3}}
(3.1) In order to show that (\ref{eq18}) cannot be uniformly strictly
positive, it is sufficient to show
\begin{equation}
\label{eq19} P \Biggl[\alpha_0(X-p_X)+\sum
_{j=1}^\infty\alpha_jH_j(X-p_j)
\Biggr]=0.
\end{equation}
Of course,
\[
P \Biggl[\alpha_0(X-p_X)+\sum
_{j=1}^\infty\alpha_jH_j(X-p_j)
\Biggr]= P\bigl[\alpha_0(X-p_X)\bigr]+P \Biggl[\sum
_{j=1}^\infty\alpha _jH_j(X-p_j)
\Biggr].
\]
Trivially,
\begin{equation}
\label{eq20} P\bigl[\alpha_0(X-p_X)\bigr]=0.
\end{equation}
Since $P$ satisfies the law of total previsions in $\pi$,
\[
P \Biggl[\sum_{j=1}^\infty
\alpha_jH_j(X-p_j) \Biggr]=P \Biggl[
P \Biggl[ \sum_{j=1}^\infty
\alpha_jH_j(X-p_i) \Big\rrvert \pi \Biggr]
\Biggr].
\]
For each $i$, $\sum_{j\ne i}\alpha_jH_j(\omega)[X(\omega)-p_j]=0$ for
all $\omega\in H_i$. It follows that, for every $i$,
\[
P \Biggl[\sum_{j=1}^\infty
\alpha_jH_j(X-p_j)\Big\rrvert H_i
\Biggr] =P\bigl[\alpha_iH_i(X-p_i)\mid
H_i\bigr],
\]
and trivially, $P[\alpha_iH_i(X-p_i)\mid H_i]=0$.
Thus,
\[
P \Biggl[\sum_{j=1}^\infty
\alpha_jH_j(X-p_i)\Big\rrvert \pi \Biggr]=0
\]
for all $\omega$, and it follows by the law of total previsions that
\begin{equation}
\label{eq23} P \Biggl[\sum_{j=1}^\infty
\alpha_jH_j(X-p_j) \Biggr] = 0.
\end{equation}
Equations (\ref{eq20}) and (\ref{eq23}) establish (\ref{eq19}).
(3.2) We must establish that there is no rival set of forecasts
$q_X$, and $\{q_j\}_{j=1}^\infty$ whose total score
uniformly dominates (\ref{eq24}). That is, there is no rival set of
forecasts such that for some $\varepsilon> 0$ and every $\omega$,
\begin{eqnarray*}
&&g_0\bigl(X(\omega),p_X\bigr)+\sum
_{j=1}^\infty H_j(\omega)g_j
\bigl(X(\omega),p_j\bigr) \\
&&\qquad\geq g_0\bigl(X(
\omega),q_X\bigr)+\sum_{j=1}^\infty
H_j(\omega)g_j\bigl(X(\omega),q_j\bigr) +
\varepsilon.
\end{eqnarray*}
It is sufficient to show that
\begin{eqnarray}
\label{eq26} \qquad &&P \Biggl\{g_0(X,q_X)+\sum
_{j=1}^\infty H_jg_j(X,q_j)
- \Biggl[g_0(X,p_X)+ \sum_{j=1}^\infty
H_jg_j(X,p_j) \Biggr] \Biggr\}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad\geq0.
\end{eqnarray}
Write the left-hand side of (\ref{eq26}) as
\begin{equation}
\label{eq27} \qquad P\bigl[g_0(X,q_X)-g_0(X,p_X)
\bigr]+P \Biggl[\sum_{j=1}^\infty
H_jg_j(X,q_j)-\sum
_{j=1}^\infty H_jg_j(X,p_j)
\Biggr].
\end{equation}
That the first expectation in (\ref{eq27}) is nonnegative follows from
the fact that $g$ is strictly proper.
From the assumption that $P$ satisfies the law of total previsions in~$\pi$,
\begin{eqnarray*}
&& P \Biggl[\sum_{j=1}^\infty
H_jg_j(X,q_j)-\sum
_{j=1}^\infty H_jg_j(X,p_j)
\Biggr]\\
&&\qquad =P \Biggl[P \Biggl[\sum_{j=1}^\infty
H_jg_j(X,q_j)-\sum
_{j=1}^\infty H_jg_j(X,p_i)
\Big\rrvert \pi \Biggr] \Biggr].
\end{eqnarray*}
Using equation (\ref{eq:diff}) and the same logic as in part (3.1), we obtain, for each~$i$,
\begin{eqnarray*}
&& P \Biggl[\sum_{j=1}^\infty
H_jg_j(X,q_j)-\sum
_{j=1}^\infty H_jg_j(X,p_j)
\Big\rrvert H_i \Biggr]\\
&&\qquad=P\bigl[H_i \bigl\{
g_i(X,q_i)-g_i(X,p_i) \bigr
\}|H_i\bigr]
\\
&&\qquad=P\bigl[g_i(X,q_i)|H_i\bigr]-P
\bigl[g_i(X,p_i)|H_i\bigr]
\\
&&\qquad\geq 0,
\end{eqnarray*}
where the final inequality follows because $g_i$ is a proper scoring
rule and $P(\cdot|H_i)$ is a finitely additive expectation for all $i$.
Therefore, since $P$ satisfies the law of total previsions in $\pi$,
\[
P \Biggl[\sum_{j=1}^\infty
H_jg_j(X,q_i)-\sum
_{j=1}^\infty H_jg_j(X,p_i)
\Biggr]\geq0,
\]
which completes the proof of (\ref{eq27}).
Corollary~\ref{thm:3} is equivalent to the claim that for each set of
rival forecasts, $q_X$ and $\{q_j\}_{j=1}^\infty$, the second
prevision in (\ref{eq27}) is nonnegative, which was established
in the proof of (3.2).
\end{appendix}
\section*{Acknowledgments}
We thank Raphael
Stern, Department of Statistics, Carnegie Mellon University,
and two anonymous referees for
helpful advice about this paper.
\begin{supplement}[id=suppA]
\stitle{Infinite previsions and finitely additive expectations\\}
\slink[doi]{10.1214/14-AOS1203SUPP}
\sdatatype{.pdf}
\sfilename{aos1203\_supp.pdf}
\sdescription{The expectation of a random variable $X$ defined on
$\Omega$ is
usually defined as the integral of $X$ over the set $\Omega$ with
respect to the underlying probability measure defined on subsets of
$\Omega$. In the countably additive setting, such integrals can be
defined (except for certain cases involving $\infty-\infty$)
uniquely from a probability measure on $\Omega$.
Dunford and Schwartz [(\citeyear{dunford-schwartz1958}), Chapter III] give a detailed analysis of
integration with respect to finitely additive measures that attempts to
replicate the uniqueness of integrals. Their analysis requires
additional assumptions if one wishes to integrate
unbounded random variables.
We choose the alternative of defining integrals as special types of
linear functionals. This is the approach
used in the study of the Daniell integral.
[See Royden (\citeyear{royden1968}), Chapter~13.] Then the measure of a set becomes the integral of its
indicator function. De Finetti's concept of
prevision turns out to be a finitely additive generalization of the
Daniell integral. (See Definition~\ref{def:daniell} in
Appendix \ref{sec:prevexp}.)
We provide details on
the finitely additive Daniell integral along with details about the
meaning of infinite previsions and how to extend coherence$_1$\vadjust{\goodbreak} and
coherence$_3$ to deal with random variables having infinite
previsions. Infinite previsions invariably arise when dealing with
general sets of unbounded random variables.}
\end{supplement}
|
\section{Integral expression of finite multiple harmonic sums}
We begin with the finite multiple harmonic sums
\[s_\mathbf{k}(N)=\sum_{N=m_1\geq\cdots\geq m_n\geq 1}
\frac{1}{m_1^{k_1}\cdots m_n^{k_n}}, \]
where $\mathbf{k}=(k_1,\ldots,k_n)$ is an $n$-tuple of positive integers
and $N$ is a positive integer.
When $k_1\geq 2$, by definition, their sum gives
the multiple zeta-star values (MZSVs for short):
\begin{equation}\label{eq:MZSV_finite}
\zeta^\star(\mathbf{k})
=\sum_{m_1\geq\cdots\geq m_n\geq 1}\frac{1}{m_1^{k_1}\cdots m_n^{k_n}}
=\sum_{N=1}^\infty s_\mathbf{k}(N).
\end{equation}
One of the basic properties of these finite multiple sums
is the following relation, called the duality:
\begin{thm}[\cite{H,K}]\label{thm:duality}
For any index $\mathbf{k}\in(\mathbb{Z}_{\geq 1})^n$ and $N\geq 0$, we have
\begin{equation}\label{eq:duality}
\sum_{i=0}^{N-1}(-1)^i\binom{N-1}{i} s_\mathbf{k}(i+1)=s_{\mathbf{k}^*}(N).
\end{equation}
Here $\mathbf{k}^*$ denotes the `transpose' of $\mathbf{k}$ (see below).
\end{thm}
For an index $\mathbf{k}=(k_1,\ldots,k_n)$, we set
\[\lvert\mathbf{k}\rvert:=k_1+\cdots+k_n, \quad
A(\mathbf{k}):=\{k_1,k_1+k_2,\ldots,k_1+\cdots+k_{n-1}\}.\]
Then the transpose $\mathbf{k}^*$ is the index determined by the property
\[\lvert\mathbf{k}\rvert=\lvert\mathbf{k}^*\rvert,\quad \{1,\ldots,\lvert\mathbf{k}\rvert-1\}
=A(\mathbf{k})\amalg A(\mathbf{k}^*). \]
For example, the transpose of $(2,3)$ is $(1,2,1,1)$.
It can be illustrated by the following picture:
\[\begin{xy}
{(0,0) \ar @{o-o} (10,0)},
{(10,0) \ar @{-o} (10,-10)},
{(10,-10) \ar @{-o} (20,-10)},
{(20,-10) \ar @{-o} (30,-10)},
{(45,0)*+{2} \ar (35,0)},
{(45,-10)*+{3} \ar (35,-10)},
{(0,-25)*+{1} \ar (0,-15)},
{(10,-25)*+{2} \ar (10,-15)},
{(20,-25)*+{1} \ar (20,-15)},
{(30,-25)*+{1} \ar (30,-15)},
\end{xy}\]
The identity \eqref{eq:duality} is somewhat analogous to
the well-known duality
\begin{equation}\label{eq:duality_MZV}
\zeta(a_1+1,\underbrace{1,\ldots,1}_{b_1-1},\ldots,
a_s+1,\underbrace{1,\ldots,1}_{b_s-1})
=\zeta(b_s+1,\underbrace{1,\ldots,1}_{a_s-1},\ldots,
b_1+1,\underbrace{1,\ldots,1}_{a_1-1})
\end{equation}
of the multiple zeta values (MZVs)
\[\zeta(\mathbf{k})=\sum_{m_1>\cdots>m_n>0}\frac{1}{m_1^{k_1}\cdots m_n^{k_n}}. \]
Since the latter duality follows immediately from
the iterated integral expression
\begin{equation}\label{eq:MZV}
\begin{split}
&\zeta(a_1+1,\underbrace{1,\ldots,1}_{b_1-1},\ldots,
a_s+1,\underbrace{1,\ldots,1}_{b_s-1})\\
&=\int_0^1\underbrace{\frac{dt}{t}\circ\cdots\circ\frac{dt}{t}}_{a_1}
\circ\underbrace{\frac{dt}{1-t}\circ\cdots\circ\frac{dt}{1-t}}_{b_1}
\circ\cdots\circ
\underbrace{\frac{dt}{t}\circ\cdots\circ\frac{dt}{t}}_{a_s}
\circ\underbrace{\frac{dt}{1-t}\circ\cdots\circ\frac{dt}{1-t}}_{b_s},
\end{split}
\end{equation}
it is natural to ask for a similar integral expression of
finite multiple sum $s_\mathbf{k}(N)$ from which \eqref{eq:duality} follows.
Here is an answer:
\begin{thm}\label{thm:finite}
Let $\mathbf{k}=(k_1,\ldots,k_n)$ be an index, and put $k=\abs{\mathbf{k}}=k_1+\cdots+k_n$.
Moreover, define
\begin{align*}
J(\mathbf{k})&=\{0,k_1,k_1+k_2,\ldots,k_1+\cdots+k_{n-1}\}=A(\mathbf{k})\cup\{0\},\\
\Delta(\mathbf{k})&=\Biggl\{(t_1,\ldots,t_k)\in[0,1]^k\Biggm|
\begin{array}{l}
t_j<t_{j+1} \text{ if $j\notin J(\mathbf{k})$},\\
t_j>t_{j+1} \text{ if $j\in J(\mathbf{k})$}
\end{array}\Biggr\}.
\end{align*}
Then, for $N\geq 1$, we have
\begin{equation}\label{eq:finite}
s_\mathbf{k}(N)=\int_{\Delta(\mathbf{k})}t_1^{N-1}dt_1
\omega_{\delta(2)}(t_2)\cdots\omega_{\delta(k)}(t_k),
\end{equation}
where $\omega_0(t)=\frac{dt}{t}$, $\omega_1(t)=\frac{dt}{1-t}$ and
\begin{equation}\label{eq:delta}
\delta(j)=\begin{cases}
0 & (j-1\notin J(\mathbf{k})), \\
1 & (j-1\in J(\mathbf{k})).
\end{cases}
\end{equation}
\end{thm}
\begin{rem}
We include $0$ in the set $J(\mathbf{k})$ to set $\delta(1)=1$ in \eqref{eq:delta},
though this value is not used in the above theorem.
We need it in Corollary \ref{cor:MZSV}.
\end{rem}
\begin{proof}
We consider the case $\mathbf{k}=(2,1,2)$ as an example,
and then the general pattern will be understood.
In this case, we should prove
\[s_{(2,1,2)}(N)=\int_{t_1<t_2>t_3>t_4<t_5}t_1^{N-1}dt_1\frac{dt_2}{t_2}
\frac{dt_3}{1-t_3}\frac{dt_4}{1-t_4}\frac{dt_5}{t_5}. \]
(Here, implicitly, the inequalities $0\leq t_i\leq 1$ are assumed.)
The right-hand side is computed by repeating single integrals, i.e.,
\begin{align*}
\int_0^{t_2}t_1^{N-1}dt_1&=\frac{t_2^N}{N},\\
\frac{1}{N}\int_{t_3}^1t_2^N\frac{dt_2}{t_2}&=\frac{1-t_3^N}{N^2},\\
\frac{1}{N^2}\int_{t_4}^1(1-t_3^N)\frac{dt_3}{1-t_3}
&=\sum_{N\geq m\geq 1}\frac{1-t_4^m}{N^2m},\\
\sum_{N\geq m\geq 1}\frac{1}{N^2m}\int_0^{t_5}(1-t_4^m)\frac{dt_4}{1-t_4}
&=\sum_{N\geq m\geq l\geq 1}\frac{t_5^l}{N^2ml},\\
\sum_{N\geq m\geq l\geq 1}\frac{1}{N^2ml}\int_0^1t_5^l\frac{dt_5}{t_5}
&=\sum_{N\geq m\geq l\geq 1}\frac{1}{N^2ml^2}.
\end{align*}
The last sum is exactly $s_{(2,1,2)}(N)$.
\end{proof}
To deduce Theorem \ref{thm:duality} from Theorem \ref{thm:finite},
note that there is a bijection
\begin{equation}\label{eq:s=1-t}
\Delta(\mathbf{k})\ni(t_1,\ldots,t_k)\longmapsto(1-t_1,\ldots,1-t_k)
\in\Delta(\mathbf{k}^*)
\end{equation}
and that the map $\delta^*$ associated with $\mathbf{k}^*$ as in \eqref{eq:delta}
satisfies $\delta^*(j)=1-\delta(j)$ for $j=2,\ldots,k$.
Hence, by changing of the integral variables $s_j=1-t_j$, we obtain
\begin{align*}
s_{\mathbf{k}^*}(N)
&=\int_{\Delta(\mathbf{k}^*)}s_1^{N-1}ds_1\,\omega_{\delta^*(2)}(s_2)
\cdots\omega_{\delta^*(k)}(s_k)\\
&=\int_{\Delta(\mathbf{k})}(1-t_1)^{N-1}dt_1\omega_{\delta(2)}(t_2)
\cdots\omega_{\delta(k)}(t_k)\\
&=\sum_{i=0}^{N-1}(-1)^i\binom{N-1}{i}
\int_{\Delta(\mathbf{k})}t_1^idt_1\omega_{\delta(2)}(t_2)
\cdots\omega_{\delta(k)}(t_k)\\
&=\sum_{i=0}^{N-1}(-1)^i\binom{N-1}{i}s_{\mathbf{k}}(i+1).
\end{align*}
By \eqref{eq:MZSV_finite} and \eqref{eq:finite},
we also obtain an integral expression of MZSVs.
\begin{crl}\label{cor:MZSV}
In the same notation as in Theorem \ref{thm:finite},
assume $k_1\geq 2$. Then
\begin{equation}\label{eq:MZSV}
\zeta^\star(\mathbf{k})=\int_{\Delta(\mathbf{k})}
\omega_{\delta(1)}(t_1)\cdots\omega_{\delta(k)}(t_k).
\end{equation}
\end{crl}
\begin{exa}
For $\mathbf{k}=(2,1)$, we have
\begin{equation}\label{eq:MZSV(2,1)}
\zeta^\star(2,1)=\int_{t_1<t_2>t_3}
\frac{dt_1}{1-t_1}\frac{dt_2}{t_2}\frac{dt_3}{1-t_3}.
\end{equation}
This integral can be divided into two parts:
\[\int_{t_1<t_2>t_3}\frac{dt_1}{1-t_1}\frac{dt_2}{t_2}\frac{dt_3}{1-t_3}
=\biggl(\int_{t_1<t_3<t_2}+\int_{t_3<t_1<t_2}\biggr)
\frac{dt_1}{1-t_1}\frac{dt_2}{t_2}\frac{dt_3}{1-t_3}. \]
By the iterated integral expression \eqref{eq:MZV} of MZVs,
the right-hand side is equal to $\zeta(2,1)+\zeta(2,1)$.
Therefore, from the integral expression \eqref{eq:MZSV(2,1)},
one obtains
\[\zeta^\star(2,1)=2\zeta(2,1). \]
Note that this is different from the relation
\[\zeta^\star(2,1)=\zeta(2,1)+\zeta(3) \]
obtained from the series expressions.
By comparing these two relations,
one proves Euler's famous relation $\zeta(2,1)=\zeta(3)$.
More generally,
from the integral and series expressions of $\zeta^\star(k-1,1)$,
one can show the sum formula for double zeta values
\[\zeta(k-1,1)+\zeta(k-2,2)+\cdots+\zeta(2,k-2)=\zeta(k)
\quad (k\geq 3)\]
in a similar manner.
\end{exa}
\section{Multiple integrals associated with 2-labeled finite posets}
Now we define a class of integrals
which includes both MZVs \eqref{eq:MZV} and MZSVs \eqref{eq:MZSV}.
Recall that a finite poset is a finite set endowed with a partial order.
In the following, we omit the word `finite' since we only consider
finite posets.
\begin{dfn}
\begin{enumerate}
\item A \emph{2-labeled poset} is a pair $X=(X,\delta_X)$ consisting of
a poset $X$ and a map $\delta_X\colon X\to\{0,1\}$,
called the \emph{labeling map}.
The \emph{weight}, denoted by $\abs{X}$,
is the number of elements of the underlying set $X$,
and the \emph{depth}, denoted by $\dep(X)$,
is the number of $x\in X$ such that $\delta(x)=1$.
\item A 2-labeled poset $X$ is said \emph{admissible}
if $\delta_X(x)=1$ for all minimal $x\in X$ and
$\delta_X(x)=0$ for all maximal $x\in X$.
\item For any poset $X$, we put
\[\Delta(X):=\bigl\{(t_x)_{x\in X}\in[0,1]^X\bigm|
t_x<t_y \text{ if } x<y\bigr\}. \]
\item For an admissible 2-labeled poset $X$,
we define the associated integral by
\begin{equation}\label{eq:I}
I(X):=\int_{\Delta(X)}\prod_{x\in X}\omega_{\delta_X(x)}(t_x).
\end{equation}
Here $\omega_0(t)=\frac{dt}{t}$ and $\omega_1(t)=\frac{dt}{1-t}$ are
the same notation as in Theorem \ref{thm:finite}.
\end{enumerate}
\end{dfn}
\begin{rem}
For the empty 2-labeled poset, denoted $\varnothing$,
we put $I(\varnothing)=1$. This is compatible with the usual definition
$\zeta(\varnothing)=\zeta^\star(\varnothing)=1$,
where $\varnothing$ denotes the index of length $0$.
\end{rem}
We use Hasse diagrams to indicate 2-labeled posets,
with vertices $\circ$ and $\bullet$ corresponding to $\delta(x)=0$ and $1$,
respectively. For example,
\[X=\{1<2<3<4<5\}\text{ and }
\bigl(\delta(1),\ldots,\delta(5)\bigr)=(1,0,1,0,0)\]
is represented as the diagram
\[\begin{xy}
{(0,8) \ar @{o-o} (0,4)},
{(0,4) \ar @{-{*}} (0,0)},
{(0,0) \ar @{-o} (0,-4)},
{(0,-4) \ar @{-{*}} (0,-8)}
\end{xy} \]
This 2-labeled poset is admissible, and we have
\[I(X)=\int_{t_1<t_2<t_3<t_4<t_5}
\frac{dt_1}{1-t_1}\frac{dt_2}{t_2}\frac{dt_3}{1-t_3}
\frac{dt_4}{t_4}\frac{dt_5}{t_5}
=\zeta(3,2). \]
In general, the iterated integral expression of a MZV
is equal to the integral $I(X)$
associated with an admissible 2-labeled \emph{totally ordered} set $X$,
and the weight and the depth of the MZV coincide with those of $X$.
Another example: Corollary \ref{cor:MZSV} for $\mathbf{k}=(2,3)$ gives
\[\zeta^\star(2,3)=I\left(\
\begin{xy}
{(0,-4) \ar @{{*}-o} (4,0)},
{(4,0) \ar @{-{*}} (8,-4)},
{(8,-4) \ar @{-o} (12,0)},
{(12,0) \ar @{-o} (16,4)}
\end{xy}\ \right). \]
Here we collect some basic constructions on 2-labeled posets,
and relations between the associated integrals.
\begin{dfn}
\begin{enumerate}
\item For 2-labeled posets $X$ and $Y$,
one can naturally define their direct sum $X\amalg Y$:
Its underlying poset is the direct sum of finite sets $X$ and $Y$
endowed with the partial order
\begin{align*}
x\leq y \text{ in $X\amalg Y$}\iff
&x,y\in X\text{ and }x\leq y\text{ in $X$ or }\\
&x,y\in Y\text{ and }x\leq y\text{ in $Y$.}
\end{align*}
The map $\delta_{X\amalg Y}\colon X\amalg Y\to\{0,1\}$
is the direct sum of the maps $\delta_X\colon X\to\{0,1\}$ and
$\delta_Y\colon Y\to\{0,1\}$.
\item Let $X=(X,\leq)$ be a poset, and $a,b\in X$ not comparable, i.e.,
neither $a\leq b$ nor $b\leq a$ hold.
Then we denote by $X_a^b$ the poset with the same underlying set $X$
and endowed with the order $\leq_a^b$ defined by
\[x\leq_a^b y\iff
x\leq y,\text{ or } x\leq a\text{ and }b\leq y. \]
We call $X_a^b$ the refinement of $X$ obtained by imposing $a<b$.
If $X$ is a 2-labeled poset, then $X_a^b$ also becomes
a 2-labeled poset with the same labeling map.
\item For a 2-labeled poset $X$, we define its \emph{transpose} $X^*$
as the 2-labeled poset consisting of the same underlying set as $X$
endowed with the reversed order (i.e.\ $x\leq y$ in $X^*$ if and only if
$y\leq x$ in $X$), and the labeling map $\delta_{X^*}(x)=1-\delta_X(x)$.
\end{enumerate}
\end{dfn}
\begin{prp}
\begin{enumerate}
\item If $X$ and $Y$ are admissible 2-labeled posets,
then $X\amalg Y$ is admissible and
\begin{equation}\label{eq:prod}
I(X\amalg Y)=I(X)\cdot I(Y).
\end{equation}
\item If $X$ is an admissible 2-labeled poset, and $a$ and $b\in X$ are
not comparable, then both $X_a^b$ and $X_b^a$ are admissible and
\begin{equation}\label{eq:sum}
I(X)=I(X_a^b)+I(X_b^a).
\end{equation}
\item If $X$ is an admissible 2-labeled poset,
then the transpose $X^*$ is admissible and
\begin{equation}\label{eq:dual}
I(X^*)=I(X).
\end{equation}
\end{enumerate}
\end{prp}
\begin{proof}
All assertions are easily verified.
Note that \eqref{eq:dual} is shown by making the change of variables
\[\Delta(X)\ni(t_x)\longmapsto (1-t_x)\in \Delta(X^*), \]
which is a generalization of \eqref{eq:s=1-t}.
\end{proof}
\begin{rem}
The shuffle relation for the MZVs can be derived from
the identities \eqref{eq:prod} and \eqref{eq:sum}
(see also the remark after Corollary \ref{cor:sum of MZVs}).
On the other hand, the identity \eqref{eq:dual} is a natural generalization
of the duality \eqref{eq:duality_MZV} for the MZVs.
\end{rem}
\begin{crl}\label{cor:sum of MZVs}
For any 2-labeled poset $X$, the integral $I(X)$ can be expressed
as the sum of a finite number of MZVs of weight $\abs{X}$
and depth $\dep(X)$.
\end{crl}
\begin{proof}
By using \eqref{eq:sum} several times, we express $I(X)$
as a sum of the integrals associated with 2-labeled totally ordered sets,
i.e., the integral expressions of MZVs.
Each of these 2-labeled totally ordered sets consists of
the same underlying set and labeling map as $X$ and
a total order extending the partial order of $X$.
In particular, these have the same weight and depth as $X$.
\end{proof}
\begin{rem}
There is an algebraic formalism for the integrals $I(X)$,
similar to the well-known pair of
the shuffle algebra $\mathfrak{H}=\mathbb{Q}\langle x,y\rangle$
and the homomorphism $Z\colon\mathfrak{H}^0=\mathbb{Q}\oplus x\mathfrak{H} y\to\mathbb{R}$.
We write $\mathfrak{P}$ for the $\mathbb{Q}$-vector space generated by
all isomorphism classes of 2-labeled posets,
and define a product on it by $[X]\cdot[Y]=[X\amalg Y]$.
Then, the subspace $\mathfrak{P}^0$ generated by admissible 2-labeled posets
is a subalgebra, and the map $I\colon \mathfrak{P}^0\to\mathbb{R}$,
defined by linearity, is indeed a $\mathbb{Q}$-algebra homomorphism.
In fact, there exists a surjective $\mathbb{Q}$-algebra homomorphism
$\rho\colon\mathfrak{P}\to\mathfrak{H}$ satisfying
$\rho(\mathfrak{P}^0)=\mathfrak{H}^0$ and $Z\circ\rho=I$.
This is defined as the unique homomorphism
whose kernel is generated by $[X]-[X_a^b]-[X_b^a]$
and which sends totally ordered $[X]$ to a monomial in $\mathfrak{H}$
encoding $\delta_X$ appropriately.
\end{rem}
\section{Other examples}
In this section, we consider some values representable
by the integrals $I(X)$, other than MZVs and MZSVs.
\subsection{Arakawa-Kaneko zeta values}
The Arakawa-Kaneko multiple zeta function \cite{AK} is defined as
\[\xi_k(s)=\frac{1}{\Gamma(s)}\int_0^\infty
\frac{Li_k(1-e^{-t})}{1-e^{-t}}e^{-t}t^{s-1}dt\]
for an integer $k>0$, where $Li_k(x)=\sum_{n=1}^\infty\frac{x^n}{n^k}$
is the $k$-th polylogarithm function.
By making the variable change $x=1-e^{-t}$, we can write it as
\[\xi_k(s)=\frac{1}{\Gamma(s)}\int_0^1
Li_k(x)\bigl(-\log(1-x)\bigr)^{s-1}\frac{dx}{x}. \]
Now we use integral expressions
\[Li_k(x)=\int_{x>t_1>\cdots>t_k>0}
\frac{dt_1}{t_1}\cdots\frac{dt_{k-1}}{t_{k-1}}\frac{dt_k}{1-t_k} \]
and
\[-\log(1-x)=\int_0^x\frac{du}{1-u}, \]
to deduce for any integer $n>0$
\begin{equation}\label{eq:AK}
\xi_k(n)=\frac{1}{(n-1)!}\,I\left(\xybox{
{(0,-9) \ar @{{*}-o} (0,-4)},
{(0,-4) \ar @{.o} (0,4)},
{(0,4) \ar @{-o} (10,9)},
{(10,9) \ar @{-{*}} (6,-5)},
{(10,9) \ar @{-{*}} (16,-5)},
{(8,-5) \ar @{.} (14,-5)},
{(-1,-9) \ar @/^1mm/ @{-} ^k (-1,4)},
{(6,-6) \ar @/_1mm/ @{-} _{n-1} (16,-6)},
}\ \right).
\end{equation}
Moreover, there are exactly $(n-1)!$ ways to impose
a total order on the $n-1$ black vertices.
Thus we have the identity
\[\xi_k(n)=I\left(\xybox{
{(0,-9) \ar @{{*}-o} (0,-4)},
{(0,-4) \ar @{.o} (0,4)},
{(0,4) \ar @{-o} (5,9)},
{(10,-9) \ar @{{*}-{*}} (10,-4)},
{(10,-4) \ar @{.{*}} (10,4)},
{(10,4) \ar @{-} (5,9)},
{(-1,-9) \ar @/^1mm/ @{-} ^k (-1,4)},
{(11,-9) \ar @/_1mm/ @{-} _{n-1} (11,4)},
}\ \right).
\]
Therefore, by \eqref{eq:MZSV}, we obtain Ohno's relation \cite[Theorem 2]{O}
\[\xi_k(n)=\zeta^\star(k+1,\underbrace{1,\ldots,1}_{n-1}). \]
\subsection{Mordell-Tornheim zeta values}
Next, we consider the values of the Mordell-Tornheim multiple zeta functions
\cite{M}
\[\zeta_{MT,r}(s_1,\ldots,s_r;s)=\sum_{m_1,\ldots,m_r>0}
\frac{1}{m_1^{s_1}\cdots m_r^{s_r}(m_1+\cdots+m_r)^s}. \]
For positive integers $k_1,\ldots,k_r,k$, it is easy to show that
\begin{equation}\label{eq:MT}
\zeta_{MT,r}(k_1,\ldots,k_r;k)=I\left(\xybox{
{(-8,-15) \ar @{{*}-o} (-8,-10)},
{(-8,-10) \ar @{.o} (-8,-2)},
{(-8,-2) \ar @{-o} (0,2)},
{(8,-15) \ar @{{*}-o} (8,-10)},
{(8,-10) \ar @{.o} (8,-2)},
{(8,-2) \ar @{-o} (0,2)},
{(0,2) \ar @{-o} (0,7)},
{(0,7) \ar @{.o} (0,15)},
{(-5,-5) \ar @{.} (5,-5)},
{(-9,-15) \ar @/^1mm/ @{-} ^{k_1} (-9,-2)},
{(9,-15) \ar @/_1mm/ @{-} _{k_r} (9,-2)},
{(1,2) \ar @/_1mm/ @{-} _k (1,15)}
}\right).
\end{equation}
For example,
\begin{align*}
I\left(\ \begin{xy}
{(-4,-4) \ar @{{*}-} (0,0)},
{(4,-4) \ar @{{*}-} (0,0)},
{(0,0) \ar @{o-o} (0,4)}\end{xy}\ \right)
&=\int_{1>t_1>t_2>0}\frac{dt_1}{t_1}\frac{dt_2}{t_2}
\biggl(\int_0^{t_2}\frac{du}{1-u}\biggr)
\biggl(\int_0^{t_2}\frac{dv}{1-v}\biggr)\\
&=\int_0^1\frac{dt_1}{t_1}\int_0^{t_1}\frac{dt_2}{t_2}
\Biggl(\sum_{m>0}\frac{t_2^m}{m}\Biggr)
\Biggl(\sum_{n>0}\frac{t_2^n}{n}\Biggr)\\
&=\sum_{m,n>0}\frac{1}{mn}
\int_0^1\frac{dt_1}{t_1}\int_0^{t_1}t_2^{m+n}\frac{dt_2}{t_2}\\
&=\sum_{m,n>0}\frac{1}{mn(m+n)}\int_0^1t_1^{m+n}\frac{dt_1}{t_1}\\
&=\sum_{m,n>0}\frac{1}{mn(m+n)^2}=\zeta_{MT,2}(1,1;2).
\end{align*}
The identity \eqref{eq:MT} implies, in particular,
the result by Bradley-Zhou \cite{BZ}
that the Mordell-Tornheim zeta value $\zeta_{MT,r}(k_1,\ldots,k_r;k)$ is
expressed as a finite sum of MZVs of weight $k_1+\cdots+k_r+k$ and depth $r$.
\subsection{Certain zeta values of root systems of type $A$}
The third class of examples is a certain type of special values
of zeta functions of root systems of type $A_N$,
considered by Komori, Matsumoto and Tsumura \cite{KMT}
in a study of shuffle relations of MZVs.
Explicitly, these values are written as
\[\zeta\biggl(p_1,\ldots,p_a;\begin{array}{c}
q_1,\ldots,q_b\\ r_1,\ldots,r_c\end{array}\biggr)
=\sum_{\substack{l_1>\cdots>l_a>m_1+n_1\\ m_1>\cdots>m_b>0\\
n_1>\cdots>n_c>0}}
\frac{1}{l_1^{p_1}\cdots l_a^{p_a}m_1^{q_1}\cdots m_b^{q_b}
n_1^{r_1}\cdots n_c^{r_c}}, \]
for three sequences $(p_1,\ldots,p_a)$, $(q_1,\ldots,q_b)$ and
$(r_1,\ldots,r_c)$ of positive integers.
To describe the corresponding diagram, we introduce an abbreviation:
For a sequence $\mathbf{k}=(k_1,\ldots,k_n)$ of positive integers,
we write
\[\begin{xy}
{(0,-3) \ar @{{*}.o} (0,3)},
{(1,-3) \ar @/_1mm/ @{-} _{\mathbf{k}} (1,3)}
\end{xy}\]
for the vertical diagram
\[\begin{xy}
{(0,-24) \ar @{{*}-o} (0,-20)},
{(0,-20) \ar @{.o} (0,-14)},
{(0,-14) \ar @{-} (0,-10)},
{(0,-10) \ar @{.} (0,-4)},
{(0,-4) \ar @{-{*}} (0,0)},
{(0,0) \ar @{-o} (0,4)},
{(0,4) \ar @{.o} (0,10)},
{(0,10) \ar @{-{*}} (0,14)},
{(0,14) \ar @{-o} (0,18)},
{(0,18) \ar @{.o} (0,24)},
{(1,-24) \ar @/_1mm/ @{-} _{k_n} (1,-14)},
{(4,-3) \ar @{.} (4,-11)},
{(1,0) \ar @/_1mm/ @{-} _{k_2} (1,10)},
{(1,14) \ar @/_1mm/ @{-} _{k_1} (1,24)}
\end{xy}\]
so that
\[\zeta(\mathbf{k})=I\left(\ \begin{xy}
{(0,-3) \ar @{{*}.o} (0,3)},
{(1,-3) \ar @/_1mm/ @{-} _{\mathbf{k}} (1,3)}
\end{xy}\right). \]
Using this notation, one can verify that
\begin{equation}\label{eq:KMT}
\zeta(\mathbf{p};\mathbf{q};\mathbf{r})=I\left(\begin{xy}
{(-4,-8) \ar @{{*}.o} (-4,-2)},
{(-4,-2) \ar @{-} (0,2)},
{(4,-8) \ar @{{*}.o} (4,-2)},
{(4,-2) \ar @{-} (0,2)},
{(0,2) \ar @{{*}.o} (0,8)},
{(-5,-8) \ar @/^1mm/ @{-} ^{\mathbf{q}} (-5,-2)},
{(5,-8) \ar @/_1mm/ @{-} _{\mathbf{r}} (5,-2)},
{(1,2) \ar @/_1mm/ @{-} _{\mathbf{p}} (1,8)}
\end{xy}\right)
\end{equation}
for $\mathbf{p}=(p_1,\ldots,p_a)$, $\mathbf{q}=(q_1,\ldots,q_b)$
and $\mathbf{r}=(r_1,\ldots,r_c)$.
In \cite{KMT}, the following relation plays an important role:
\begin{equation}\label{eq:KMTrel}
\begin{split}
&\zeta\biggl(p_1,\ldots,p_a;\begin{array}{c}
q_1,\ldots,q_b\\ r_1,\ldots,r_c\end{array}\biggr)\\
&=\sum_{j=0}^{q_1-1}\binom{r_1-1+j}{j}\,
\zeta\biggl(p_1,\ldots,p_a,r_1+j;\begin{array}{c}
q_1-j,q_2,\ldots,q_b\\ r_2,\ldots,r_c\end{array}\biggr)\\
&+\sum_{j=0}^{r_1-1}\binom{q_1-1+j}{j}\,
\zeta\biggl(p_1,\ldots,p_a,q_1+j;\begin{array}{c}
q_2,\ldots,q_b\\ r_1-j,r_2,\ldots,r_c\end{array}\biggr).
\end{split}
\end{equation}
We point out that our expression \eqref{eq:KMT} implies this relation
quite naturally.
To do this, we denote by $X$ the 2-labeled poset indicated in \eqref{eq:KMT},
and name some vertices as follows:
\[\begin{xy}
{(-6,-28) \ar @{{*}.o} (-6,-20)},
{(-6,-20) \ar @{-} (-6,-14)},
{(-6,-14) \ar @{{*}.o} (-6,-2)},
{(-10,-14)*++{y_{q_1}} \ar @{.} (-10,-2)*++{y_1}},
{(-6,-2) \ar @{-} (0,2)},
{(6,-28) \ar @{{*}.o} (6,-20)},
{(6,-20) \ar @{-} (6,-14)},
{(6,-14) \ar @{{*}.o} (6,-2)},
{(10,-14)*++{z_{r_1}} \ar @{.} (10,-2)*++{z_1}},
{(6,-2) \ar @{-} (0,2)},
{(0,2)*+!U{x} \ar @{{*}.o} (0,10)},
{(-7,-28) \ar @/^1mm/ @{-} ^{\mathbf{q}'} (-7,-20)},
{(7,-28) \ar @/_1mm/ @{-} _{\mathbf{r}'} (7,-20)},
{(1,2) \ar @/_1mm/ @{-} _{\mathbf{p}} (1,10)}
\end{xy}\]
where $\mathbf{q}'=(q_2,\ldots,q_b)$ and $\mathbf{r}'=(r_2,\ldots,r_c)$.
By \eqref{eq:sum}, one has
\begin{equation}\label{eq:KMT1}
I(X)=I(X_{y_{q_1}}^{z_{r_1}})+I(X^{y_{q_1}}_{z_{r_1}}).
\end{equation}
In $X_{y_{q_1}}^{z_{r_1}}$,
the inequalities $x>z_{r_1}>y_{q_1}$ and $x>y_1>\cdots>y_{q_1}$ hold,
hence one can consider $q_1$ further refinements by imposing
$y_j>z_{r_1}>y_{j+1}$ for $j=0,\ldots,q_1-1$ (here we write $y_0=x$).
Thus the identity
\begin{equation}\label{eq:KMT2}
I(X_{y_{q_1}}^{z_{r_1}})=\sum_{j=0}^{q_1-1}I(X_j)
\end{equation}
holds, where $X_j$ is represented by the diagram
\[\begin{xy}
{(-6,-38) \ar @{{*}.o} (-6,-30)},
{(-6,-30) \ar @{-{*}} (-6,-26)},
{(-6,-26) \ar @{.o} (-6,-18)},
{(-11,-26)*+{y_{q_1}} \ar @{.} (-11,-18)*+{y_{j+1}}},
{(-6,-18) \ar @{-} (0,-14)},
{(0,-14)*+!U{z_{r_1}} \ar @{{*}-o} (-6,-10)},
{(-6,-10) \ar @{.o} (-6,-2)},
{(-11,-10)*+{y_j} \ar @{.} (-11,-2)*+{y_1}},
{(-6,-2) \ar @{-} (0,2)},
{(6,-26) \ar @{{*}.o} (6,-18)},
{(6,-18) \ar @{-} (0,-14)},
{(0,-14) \ar @{-o} (6,-10)},
{(6,-10) \ar @{.o} (6,-2)},
{(11,-10)*+{z_{r_1-1}} \ar @{.} (11,-2)*+{z_1}},
{(6,-2) \ar @{-} (0,2)},
{(0,2)*+!U{x} \ar @{{*}.o} (0,10)},
{(-7,-38) \ar @/^1mm/ @{-} ^{\mathbf{q}'} (-7,-30)},
{(7,-26) \ar @/_1mm/ @{-} _{\mathbf{r}'} (7,-18)},
{(1,2) \ar @/_1mm/ @{-} _{\mathbf{p}} (1,10)}
\end{xy}\]
Moreover, since there are $\binom{r_1-1+j}{j}$ ways to impose
a total order on the white vertices $y_1,\ldots,y_j$,
$z_1,\ldots,z_{r_1-1}$, one has
\begin{equation}\label{eq:KMT3}
I(X_j)=\binom{r_1-1+j}{j}\,
\zeta\biggl(p_1,\ldots,p_a,r_1+j;\begin{array}{c}
q_1-j,q_2,\ldots,q_b\\ r_2,\ldots,r_c\end{array}\biggr).
\end{equation}
The identities \eqref{eq:KMT2} and \eqref{eq:KMT3}
expresses the first term of \eqref{eq:KMT1}
as desired in \eqref{eq:KMTrel}.
The second is obtained in the same way.
\begin{rem}
In \cite{KMT}, using partial fraction decompositions,
the identity \eqref{eq:KMTrel} is proved with some variables
(irrelevant to the decomposition) \emph{complex valued},
not necessarily positive integral.
It seems difficult to apply our method in this paper
to such functional relations.
\end{rem}
|
\section{Introduction}
The expansion velocities in supernovae (SN) are quite high
and, as an example, a time series of eight spectra in SN\, 2009ig~
reports that the velocity at the \rm{CA}\,{\small\rmfamily II}\, line,
decreases from 32000 $\rm\,{kms^{-1}}$ to 21500$\rm\,{kms^{-1}}$,
in 12 day,
see Fig. 9 in \cite{Marion2013}.
Another example is given
by SN\, 2009bb~ in which the velocity
of expansion has been evaluated
to be
$\approx$ 255000 $\rm\,{kms^{-1}}$, see \cite{Soderberg2010}.
We briefly recall that the corrections in special relativity (SR)
for stable atomic clocks in satellites
of the
Global Positioning System (GPS) are applied to satellites which are moving
at a velocity of $\approx 3.87 \rm\,{kms^{-1}}$.
The problem of the aspherical SN, such as SN \,1987A\,,
is to find an acceptable model which can reproduce
the observed complex morphology of the aspherical SN \,1987A\,
and this was done in a classical framework by
\cite{Zaninetti2013c}.
In this paper we shall discuss a relativistic treatment
of the thin layer approximation in the presence of an auto-gravitating medium.
\section{Relativistic conservation of momentum}
The chosen auto-gravitating profile is
\begin{equation}
n(R,\theta) = n_0 sech^2 (\frac{R \sin (\theta) }{2\,h})
\quad ,
\label{sech2rtheta}
\end{equation}
where
$n_0$ is the density in the equatorial plane ($\theta=0$),
$R$ is the radius of the advancing shell,
$\theta$ is
the latitude angle ($\theta=0$ at the equator and $\theta=\pm 90$
at the two poles)
and $h$ is a parameter which characterizes the gradient.
The chosen symmetry imposes that the motion is independent of the azimuthal angle in spherical coordinates but depends only on the latitude angle and the time.
The classical conservation of momentum
in the presence of an auto-gravitating medium
was treated in \cite{Zaninetti2013c} and therefore
we will not duplicate the results already obtained.
The relativistic conservation of momentum,
see \cite{French1968,Zhang1997,Guery2010},
is formulated as
\begin{equation}
M(R_0;b) \gamma_0 \beta_0 = M(R;b) \gamma \beta
\quad ,
\label{momentumrelativistic}
\end{equation}
with
\begin{equation}
\gamma_0 = \frac{1} {
\sqrt{1-\beta_0^2}
}
\quad ; \qquad
\gamma = \frac{1} {
\sqrt{1-\beta^2}
}
\quad ,
\end {equation}
and
\begin{equation}
\beta_0 =\frac{v_0}{c}
\quad ; \qquad
\beta =\frac{v}{c}
\quad ,
\end{equation}
$c$ being the velocity of light,
here $M(R_0;b)$ is a first mass between 0 and $R_0$ and
$M(R;b)$ is a second mass between 0 and $R$.
We know already that $M(R;b) = (I_m(R))^{1/p}$
where the integral $I_m(R)$ has been defined in eq. (15) of
\cite{Zaninetti2013c} and $p$ is a parameter to be found.
The fundamental Eq. (\ref{momentumrelativistic}) can be first
solved for $\beta^2$
\begin{equation}
\beta^2 = \frac{N}{D}
\quad ,
\end{equation}
where
\begin{eqnarray}
N ={{\it R_0}}^{6\,{p}^{-1}}{{\it \beta_0}}^{2}
\quad ,
\nonumber
\end{eqnarray}
\begin{eqnarray}
D= - ( -1 ) ^{2\,{p}^{-1}} ( -C{{\it R0}}^{3}{S}^{3}+8\,C
\ln ( 1+B ) R{h}^{2}S
\nonumber \\
-8\,C\ln ( 1+A ) {\it R0}
\,{h}^{2}S+4\,C{{\it R0}}^{2}h{S}^{2}-4\,C{R}^{2}h{S}^{2}
\nonumber \\
-B{{\it R0}}^
{3}{S}^{3}-A{{\it R0}}^{3}{S}^{3}+8\,CP ( 2,-B ) {h}^{3}-8
\,CP ( 2,-A ) {h}^{3}
\nonumber \\
+8\,B\ln ( 1+B ) R{h}^{2}S
-8\,B\ln ( 1+A ) {\it R0}\,{h}^{2}S-4\,B{R}^{2}h{S}^{2}
\nonumber \\
+8
\,A\ln ( 1+B ) R{h}^{2}S-8\,A\ln ( 1+A ) {\it
R0}\,{h}^{2}S+4\,A{{\it R0}}^{2}h{S}^{2}
\nonumber \\
-{{\it R0}}^{3}{S}^{3}+8\,BP
( 2,-B ) {h}^{3}-8\,BP ( 2,-A ) {h}^{3}+8\,AP
( 2,-B ) {h}^{3}
\nonumber \\
-8\,AP ( 2,-A ) {h}^{3}+8\,{h}^
{2}R\ln ( 1+B ) S-8\,{h}^{2}{\it R0}\,\ln ( 1+A
) S
\nonumber \\
+8\,{h}^{3}P ( 2,-B ) -8\,{h}^{3}P ( 2,-A
) ) ^{2\,{p}^{-1}} ( 1+B ) ^{-2\,{p}^{-1}}{S}
^{-6\,{p}^{-1}} ( 1+A ) ^{-2\,{p}^{-1}}{{\it \beta_0}}^{2}
\nonumber \\
+
( -1 ) ^{2\,{p}^{-1}} ( -C{{\it R_0}}^{3}{S}^{3}+8\,C
\ln ( 1+B ) R{h}^{2}S-8\,C\ln ( 1+A ) {\it R_0}
\,{h}^{2}S
\nonumber \\
+4\,C{{\it R_0}}^{2}h{S}^{2}-4\,C{R}^{2}h{S}^{2}-B{{\it R_0}}^
{3}{S}^{3}-A{{\it R_0}}^{3}{S}^{3}+8\,CP ( 2,-B ) {h}^{3}
\nonumber \\
-8
\,CP ( 2,-A ) {h}^{3}+8\,B\ln ( 1+B ) R{h}^{2}S
-8\,B\ln ( 1+A ) {\it R_0}\,{h}^{2}S-4\,B{R}^{2}h{S}^{2}
\nonumber \\
+8
\,A\ln ( 1+B ) R{h}^{2}S-8\,A\ln ( 1+A ) {\it
R_0}\,{h}^{2}S+4\,A{{\it R_0}}^{2}h{S}^{2}-{{\it R_0}}^{3}{S}^{3}
\nonumber \\
+8\,BP
( 2,-B ) {h}^{3}-8\,BP ( 2,-A ) {h}^{3}+8\,AP
( 2,-B ) {h}^{3}-8\,AP ( 2,-A ) {h}^{3}
\nonumber \\
+8\,{h}^
{2}R\ln ( 1+B ) S-8\,{h}^{2}{\it R_0}\,\ln ( 1+A
) S+8\,{h}^{3}P ( 2,-B )
\nonumber \\
-8\,{h}^{3}P ( 2,-A
) ) ^{2\,{p}^{-1}} ( 1+B ) ^{-2\,{p}^{-1}}{S}
^{-6\,{p}^{-1}} ( 1+A ) ^{-2\,{p}^{-1}}+{{\it R_0}}^{6\,{p}^
{-1}}{{\it \beta_0}}^{2}
\quad ,
\nonumber
\end{eqnarray}
with
\begin{eqnarray}
A= & {{\rm e}^{{\frac {R_{{0}}\sin \left( \theta \right) }{h}}}}
\nonumber \\
B= & {{\rm e}^{{\frac {R\sin \left( \theta \right) }{h}}}}
\nonumber \\
C= & {{\rm e}^{{\frac {\sin \left( \theta \right) \left( R_{{0}}+R
\right) }{h}}}}
\nonumber \\
S= & \sin \left( \theta \right)
\nonumber
\quad .
\end{eqnarray}
and $P$ the polylog operator, which is
defined by
\begin{equation}
polylog(a,z) = \sum
_{{n=1}}^{\infty}\frac{z^{n}}{n^{a}}
\quad .
\end{equation}
The value of $\beta$ is
\begin{equation}
\beta = \sqrt{\frac{N}{D}}
\quad ,
\end{equation}
or
\begin{equation}
\frac{dR}{dt} = c \sqrt{\frac{N}{D}}
\label{eqndiff}
\quad .
\end{equation}
This first order differential equation can be solved with
the Runge--Kutta method, see
FORTRAN SUBROUTINE RK4 in \cite{press}.
Another approach
separates the variables
\begin{equation}
\int_{R_0}^R
\frac{1}{\sqrt{\frac{N}{D}}}
\,dR
= c (t-t_0)
\quad .
\label{nlequation}
\end{equation}
The previous integral does not have an analytical solution
and we treat the previous result as a non-linear equation to be
solved with the FORTRAN SUBROUTINE ZRIDDR in \cite{press}.
The presence of an analytical expression for $\beta$
as given by Eq. (\ref{eqndiff}) allows setting up the
recursive solution
\begin{eqnarray}
R_{n+1} = & R_n + V_n(R_0,R_n,\beta_0,h) \Delta t \nonumber \\
V_{n+1} = & V_n(R_0,R_{n+1},\beta_0,h)
\quad ,
\label{recursiverel}
\end{eqnarray}
where $R_n$, $V_n$, $\Delta t$ are the temporary radius,
the relativistic velocity,
and the interval of time, respectively.
An interesting application of SR is the time delay:
given an interval of time, $\Delta t$,
in the laboratory frame the interval of time,
$\Delta t^{\prime} $,
in a frame that that is moving with
velocity $v$ in the $x$-direction is
\begin{equation}
\Delta t^{\prime} = \frac{\Delta t}{\sqrt{1-\frac{v^2}{c^2}}}
\quad .
\end{equation}
We can therefore introduce the following ratio
\begin{equation}
D = \frac{\Delta t } {\Delta t^{\prime}}
\quad ,
\label{dilation}
\end{equation}
which measures the time dilation, and lies between 0 and 1.
\section{Astrophysical application}
We numerically solved the
non-linear equation, Eq. (\ref{nlequation})
even if the same results can be obtained
by solving the differential equation
(\ref{eqndiff})
or implementing the
recursive relationship
as given by
Eq. (\ref{recursiverel}), see
Table \ref{datafitrel1987a} for the
adopted data.
\begin{table}
\caption
{
The numerical values of the parameters
of the relativistic simulation for
SN \,1987A\,
}
\label{datafitrel1987a}
\[
\begin{array}{lcc}
\hline
\hline
\mbox{Quantity} & \mbox{Unit} & \mbox{value} \\
R_0 & \mbox{pc} & 0.011 \\
\dot {R}_0 & \mbox{km s}^{-1} & 30000 \\
p & \mbox{number} & 4 \\
h & \mbox{pc} & 0.01 \\
t_{0} & \mbox{yr} & 0.00022 \\
t & \mbox{yr} & 23 \\
\noalign{\smallskip}
\hline
\hline
\end{array}
\]
\end {table}
The complicated structure of SN \,1987A\,
is due to the great variety of shapes
obtained when the point
of view of the observer
changes.
One way to parametrize the point of view
of the observer
is the introduction of
the Euler angles $(\Phi, \Theta, \Psi)$,
as an example, Fig. \ref{rel_sn1987a_faces_rotated}
shows the 3D advancing shell
after 23 years.
\begin{figure}
\begin{center}
\includegraphics[width=10cm]{f01.pdf}
\end {center}
\caption {
Continuous three-dimensional surface of SN \,1987A\,
after 23 yr: the three Eulerian angles characterizing the point of view are
$ \Phi $=105$^{\circ }$,
$ \Theta $=55 $^{\circ }$
and $ \Psi $=-165 $^{\circ }$.
Physical parameters as in Table
\ref{datafitrel1987a}.
}%
\label{rel_sn1987a_faces_rotated}
\end{figure}
In order to avoid complicated changes
of framework for the field of velocity
we limit ourselves to the non-rotated image.
This choice is already widely used by
astronomers in order to reduce the data
of $\eta$-Carinae~, see Fig. 4 in \cite{Smith2006}.
The progressive increase of the asymmetry
is clearly outlined in Fig.
\ref{rel_sezioni_1987a}, in which sections
of the expansion are drawn at time steps of 1 yr.
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{f02.pdf}
\end{center}
\caption
{
Sections of SN \,1987A\, in the {\it X-Z} plane
at time steps of 1yr.
Physical parameters as in Table \ref{datafitrel1987a}.
This is a non-rotated image
and the three Euler angles
characterizing the orientation
are $ \Phi $=180$^{\circ }$,
$ \Theta $=90 $^{\circ }$
and $ \Psi $=0 $^{\circ }$.
}
\label{rel_sezioni_1987a}%
\end{figure}
The difference in velocity between
the polar direction and equatorial direction
are oulined in Fig. \ref{duevel_sn1987a}.
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{f03.pdf}
\end{center}
\caption
{
Velocity in the equatorial direction (full line)
and in the polar direction (dashed line)
in the first 50 days.
}
\label{duevel_sn1987a}%
\end{figure}
The relativistic field of velocity in the various
points of SN \,1987A\,
after 1 yr was shown
in Fig. \ref{sn1987a_v_field_rel}.
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{f04.pdf}
\end {center}
\caption
{
Map of the relativistic velocity
as a function of the latitude
for SN \,1987A\, at the age 1 yr.
}%
\label{sn1987a_v_field_rel}
\end{figure}
The relativistic time dilation is mapped in
Fig. \ref{timedilation} where
the velocity of expansion perpendicular to
the observer ($x$-direction) is considered.
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{f05.pdf}
\end {center}
\caption
{
Map of the relativistic time dilation
$D$, see Eq. (\ref{dilation}),
for velocity of SN \,1987A\, in the direction perpendicular
to the observer at the age 1 yr at time intervals
of 1 s in the laboratory frame.
}
\label{timedilation}
\end{figure}
\section{Conclusions}
We have covered the evolution of a SN in an auto-gravitating medium
in a relativistic framework. The initial shape is represented
by a sphere of radius $R_0=0.011$\ pc.
After 1 yr, the asymmetry between the radius
in the equatorial plane
and the radius in the polar direction is
well defined
and Fig. \ref{sn1987a_v_field_rel}
summarizes both the asymmetrical
shape and the anisotropic field of velocity.
The time dilation at 1 yr as represented
by the parameter $D$ varies
between a minimum of 0.9975 and a maximum
of 1.
\noindent
{\bf REFERENCES}
\providecommand{\newblock}{}
|
\section{Introduction}
Magnetoelasticity (or magnetostriction) is the property of certain solids exhibiting
a strong coupling between mechanical and magnetic variables. As effect
of this coupling, relevant reversible mechanical deformations can
be induced by the application of an external magnetic field. This
behavior is clearly of a great applicative interest in connection
with sensors and actuators design, as well as for a variety of innovative functional-material devices.
The origin of magnetoelasticity lies in the interplay between
material crystallographic patterning (where different crystals present
different easy axis of magnetization) and magnetic domains. In absence
of external magnetic fields, magnetic domains orient in such a way to
minimize long-range dipolar effects. This generically results in some
small or even negligible magnetization of the medium. Upon applying an external
magnetic field the magnetic
domains tend to reorient toward it. As magnetizations are
related to specific stress-free reference strains, this causes indeed
the emergence of a macroscopic deformation. As the intensity of the
magnetic field is increased, more and more
magnetic domains orientate themselves so that
their principal axes of anisotropy are collinear
with the magnetic field in each region and finally
saturation is reached. We refer to e.g.~\cite{brown} for a
discussion of the physical
foundations of magnetoelasticity.
The mathematical modeling of magnetoelasticity is a vibrant
area of research, in particular, in connection with intelligent
materials such as iron/rare-earth giant magnetostrictive materials \cite{DKV,james-kinderlehrer,K9} and
magnetic shape-memory alloys \cite{bks,bs,JS}. Correspondingly, the
understanding of the statics of these materials has attracted
considerable attention
\cite{DD98,desimone-james,rybka-luskin}. Building from one
side upon the static thin-film-limit analysis for
magnetic materials in {\sc Gioia \& James} \cite{goia-james} and
from the other side on the dimension reduction for static linear elastic plates
in \cite{freddi-paroni-zanini}, we shall be here concerned with the evolutive situation instead. In
particular, we are interested in a slow quasistatic evolution
of such materials under the combined action of conservative and
dissipative forces. We indeed assume that the change of
magnetization implies
dissipation while no dissipation is associated with elastic
variables. The evolution is driven by a Dirichlet boundary condition
and/or by an external magnetic field. Changes of external
conditions are considered to be slow enough so that inertial effects can be
neglected and the system is always in
equilibrium so that its evolution is quasistatic.
The focus of the paper is on deriving a quasistatic evolution
theory for magnetoelastic thin films. After considering the
above-mentioned dissipative evolution problem for the bulk,
three-dimensional material, we address the thin-film evolution by
means of a dimension reduction argument. In particular, we assume that
the reference configuration of the body is thin in one dimension and
we pass to the limit with respect to it. Our
specific choice for the scaling entails that in the limit
one obtains a Kirchhoff-Love quasistatic evolution plate model
for magnetoelastic materials. Moreover, we are able to deal
with thickness-depedent change of the anisotropic magnetic behavior
of the sample, both at the static and the evolutive level.
The novelty of the paper is indeed twofold. From the one hand, we
advance the first quasistatic evolution model for
magnetoelastic thin films and prove the existence of suitable
variational solutions. In particular, the emergence of the so-called
magnetic anisotropic behavior is emphasized. Secondly, by deriving such a model
by dimension reduction, we provide a novel evolutive
$\Gamma$-convergence result in the magnetoelastic context.
This consists in combining some slightly refined
version of the already available static thin-film-limit theory within the general
frame of the evolutive
$\Gamma$-convergence analysis for rate-independent systems from
\cite{mrs}. We shall observe that, as already in \cite{goia-james},
the magnetostatic energy contribution which is usually
difficult to evaluate in micromagnetics reduces in our model to
calculating the square of the third component of the
magnetization. This makes the model attractive from the
point of view of numerics.
Apart from the magnetoelastic setting,
dimension reduction via $\Gamma$-convergence in the quasistatic evolutive setting has already
attracted some attention.
{\sc Liero \&
Mielke} derive in \cite{ML11,M12} an elastoplastic plate theory in
presence of linear kinematic hardening. A
different theory is then obtained by an alternative scaling choice by
{\sc Liero \& Roche} \cite{LR12}. The perfectly-plastic case has then
be considered by {\sc Davoli \& Mora} \cite{DM13} and {\sc
Davoli} \cite{D13,D14}, also in the frame of finite
plasticity.
{\sc Babadjian} obtained in \cite{Ba06} via dimension reduction the existence of a quasistatic evolution for a free crack in an elastic brittle thin film.
Dimension reduction in a delamination context is addressed
in \cite{FPRZ,FRZ13} whereas an application to shape-memory thin films
is described in \cite{BKP}.
The plan of the paper is as follows. We start by describing the bulk
model in the static three-dimensional situation in Section
\ref{description}. Then, the
corresponding static thin-film micromagnetic and magnetoelastic limits are
discussed in Section~\ref{dimension-reduction}. Eventually, Section
\ref{evo} focuses on quasistatic evolution situations both in the
bulk (Subsection~\ref{analysis}) and in the thin-limit case
(Subsection \ref{KL}, respectively). In particular, Subsection
\ref{KL} contains our main convergence result, i.e. Theorem~\ref{thm}.
\section{ Description of the static bulk model}\label{description}
\setcounter{equation}{0}
Let us start by specifying the modelization in the
three-dimensional setting. The thin-film model will then be derived in
Section \ref{evo} by means of a rigorous dimension reduction procedure. We assume to have fixed
an orthonormal basis $\{e_1,e_2,e_3\}$ of ${\mathbb R}^3$ and to be given a thin
magnetic body with reference configuration
$\Omega_h:=\{(x_1,x_2)\in S;\ 0<x_3<h\}$. Here, $S\subset{\mathbb R}^2$
is a bounded Lipschitz domain in the $\{e_1,
e_2\}$-plane and $h >0 $ represents the small thickness of the
specimen, eventually bound to go to $0$.
\subsection{Micromagnetics}
The magnetization of the body is described by $\bar
m:\Omega_h\rightarrow {\mathbb R}^3$ subject to the saturation constraint
$$|\bar m(x)|=m_{\textrm{sat}} \ \text{ for a.e. } x\in\Omega_h,$$
where the {\it saturation magnetization} $m_{\textrm{sat}}>0$ is assumed to be
constant. The micromagnetic energy of the film is
classically defined as \cite{brown,desimone-james,kruzik-prohl}
\begin{align}
\bar E^{\rm mag}_h(t,\bar m)&:=\frac{1}{|S|h} \int_{\O_h}\left(\alpha
|\nabla \bar m(x)|^2{+}\bar \varphi_h (x, \bar m(x)){+}\frac12\bar m(x)\cdot\nabla
\bar \xi \right) {\rm d} x \nonumber\\
& - \int_{\O_h} \bar H(x)\cdot \bar m(x) {\rm d} x. \label{micromagenergy}
\end{align}
The {\it stray field} $\nabla \bar \xi$ is related to
the magnetization $\bar m$ via the Maxwell equation
\begin{equation}
\nabla\cdot(-\mu_0\nabla \bar \xi+\bar m\chi_{ \O_h})=0 \ \mbox{ in }{\mathbb R}^3\nonumber
\end{equation}
where $\mu_0$ is the vacuum permeability and
$\chi_{ \O_h}$ is the {\it characteristic function} of the domain
$\O_h$, namely $\chi_{ \O_h}=1$ on $\O_h$ and $\chi_{ \O_h}=0$
elsewhere in ${\mathbb R}^3$.
The first term in the integral in (\ref{micromagenergy}) is the {\it exchange
energy}, penalizing indeed spatial changes of the
magnetization.
The (thickness-dependent) {\it magnetic potential}
$\bar \varphi_h : \Omega_h \times m_{\textrm{sat}} S^2 \to [0,\infty)$ describes the
magnetic anisotropy of the material. In particular, for all
thicknesses $h>0$ and $x \in \Omega_h$ it is an even function
vanishing precisely at the set
$\{\pm s_i;\,|s_i|=m_{\textrm{sat}}\}_{i=1}^N$ for some $s_i=s_i(x)$, where
$N=1$ for uniaxial magnets and $N=3$ or $N=4$ for cubic magnets.
The lines through $\pm s_i$ are called {\it easy axes} of
the magnet. The space dependence in $\bar \varphi_h$ is intended
to model the polycrystalline texture of the medium and we assume $\bar
\varphi_h$ to be continuous. This particularly entails that the anisotropic energy
term is lower semicontinuous with respect to the $L^2$ topology.
Following the classical theory by N\'eel
\cite{Nell54}, we allow the magnetic anisotropy of the medium to depend
on the sample thickness. It is indeed observed that many material
systems develop a very strong magnetic anisotropy in the off-plane
direction as $h \to 0$, see \cite{Bruno88,Schulz94}, for instance. This effect
is at the basis of the so-called {\it perpendicular recording}
technology, see the review \cite{Richter07}. \color{black}
The term containing $\nabla \xi$ is the so-called {\it
stray-field energy} and represents long-range
dipolar self-interactions favoring indeed the formation of a
solenoidal magnetic field. In particular, $\xi$ is the {\it
magnetostatic potential}. Eventually, the last term in the
right-hand side of
\eqref{micromagenergy} is the {\it Zeeman energy}, namely
the work done by the external magnetic field $\bar H \in
L^1(\O_h;{\mathbb R}^3)$. We anticipate that in Section \ref{evo} the
external field will
depend on time and drive the quasistatic evolution of the film.
An application of the Direct Method of the Calculus of
Variations, see e.g.~\cite{james-kinderlehrer}, ensures
that for every $h>0$, the micromagnetic energy $\bar E^{\rm
mag}_h$ admits a minimizer in the set
$$\mathcal M_h:=\{\bar m\in W^{1,2}(\O_h;{\mathbb R}^3) \ : \ |\bar m | = m_{\textrm{sat}}\mbox{ a.e.}\}.$$
\subsection{Magnetomechanics}
The medium will be subject to nonhomogeneous time-dependent Dirichlet boundary
conditions on some distinguished part $\Gamma_h:= \omega \times (0,h)$ of the boundary
$\partial S \times (0,h)$ where $\omega \subset \partial S $ is of positive surface measure. In order to
prescribe these conditions we assume to be given $ \bar u^{\rm Dir}_h\in
W^{1,2}(\O_h;{\mathbb R}^3)$ and let $$\bar u + \bar u^{\rm
Dir}_h:\O_h\to{\mathbb R}^3$$ be the displacement of the specimen from its
reference configuration. We classically denote by
$\varepsilon(\bar u)$ the symmetrized gradient
$\varepsilon(\bar u):=(\nabla \bar u+\nabla^\top \bar u)/2$.
Within the small deformation realm, we linearly decompose the strain of the
material as
$$ \varepsilon(\bar u {+} \bar u^{\rm
Dir}_h) = \varepsilon^{\rm elas} + \varepsilon^{\rm mag}(\bar
m).$$
Here, $\varepsilon^{\rm elas} $ is the elastic part of the strain. In
particular, $\varepsilon^{\rm elas} = \mathbb{C}^{-1}\sigma$, where $\mathbb{C}$ is the elasticity tensor (symmetric,
positive definite) and $\sigma$ is the {\it stress} experienced by the
material. On the other hand,
$\varepsilon^{\rm mag}(\bar
m)$ is the stress-free strain corresponding to the magnetization $\bar
m$. In particular, we could choose
$$\varepsilon^{\rm mag}(\bar m):=\bar m\otimes \bar m-\frac{m_{\textrm{sat}}^2}{3}\mathbb{I},$$
where $\mathbb{I}$ is the identity matrix in ${\mathbb R}^{3\times 3}$.
Note
that $\varepsilon^{\rm mag}$ is a symmetric, continuous, even, and
deviatoric (as $|\bar m|=m_{\textrm{sat}}$) tensor-valued mapping of $\bar m$. The specific
form of $\varepsilon^{\rm mag}$ is here chosen for definiteness only. In
fact, other forms of $\varepsilon^{\rm mag}$ can also be covered by our model
as long as they enjoy the mentioned properties.
The {\it elastic energy} of the
medium is classically described by the quadratic form
\begin{align}
\bar E^{\rm elas}_h(t,\bar u,\bar
m)& :=\frac{1}{2|S|h}\int_{\O_h}\mathbb{C} \varepsilon^{\rm elas}{:}
\varepsilon^{\rm elas} \, {\rm d} x\nonumber\\
& =\frac{1}{2|S|h}\int_{\O_h}\mathbb{C}\big(\varepsilon(\bar
u{+}\bar u^{\rm Dir}_h)-\varepsilon^{\rm mag}(\bar m) \big){:}\big (\varepsilon(\bar u{+}\bar u^{\rm Dir}_h)-\varepsilon^{\rm mag}(\bar m) \big)\,{\rm d} x.\nonumber
\end{align}
Given the magnetization $\bar m$, the elastic equilibrium problem
consists in finding $\bar u$ minimizing the elastic energy
$\bar E^{\rm elas}_h$
on the set of admissible displacements
\begin{align}
\mathcal U_h:=\{u\in W^{1,2}(\O_h;{\mathbb R}^3) \ : \ u=0 \mbox{ on }\Gamma_h\}.\nonumber
\end{align}
This problem has clearly a unique solution which depends linearly
on both $\varepsilon(\bar u^{\rm Dir}_h)$ and $\varepsilon^{\rm mag}(\bar m)$. Note that, in
particular, $\bar u+\bar u^{\rm Dir}_h= \bar
u^{\rm Dir}_h$ on $\Gamma_h$.
The total magnetoelastic energy of the specimen results from the
sum of the micromagnetic and the elastic energy and reads
\begin{align}
\bar E_h(\bar u,\bar m):=\bar E^{\rm mag}_h(\bar m) +
\bar E^{\rm elas}_h(\bar u,\bar m).
\nonumber
\end{align}
It is a rather standard matter to check that the total energy
admits minimizers $(\bar u ,\bar m)$ in the set $\mathcal U_h \times
\mathcal M_h$. These correspond to a variational solution of the
magnetoelastic system
\begin{subequations}\label{sys}
\begin{align}
\nabla {\cdot} \sigma &=0 \quad \text{in} \ \Omega_h,\\
\mathbb{C}(\varepsilon(\bar u {+} \bar u^{\rm Dir}_h)-\varepsilon^{\rm
mag}(\bar m)) &=\sigma \quad \text{in} \ \Omega_h,\\
- \alpha \Delta \bar m + \nabla_{\bar m} \bar \varphi_h(x,
\bar m) +\frac12 \nabla
\bar
\xi &=\bar H \quad \text{in} \ \Omega_h,\label{m}\\
\nabla\cdot(-\mu_0\nabla \bar \xi+\bar m\chi_{ \O_h})&=0\quad
\text{in} \ {\mathbb R}^3,\\
\alpha \partial_{\nu} \bar m &=0 \quad \text{on} \ \partial \Omega_h,\\
\bar u&=0 \quad \text{on} \ \Gamma_h
\end{align}
\end{subequations}
where we have denoted by $\nu$ the outer unit normal to $\partial \Omega_h$.
\section{ Static thin-film limits }\label{dimension-reduction}
We shall preliminarily record here some dimension reduction analysis in the static situation of Section \ref{description}. Our aim is to investigate the limit $h \to 0$,
corresponding indeed to the situation of a very thin structure in the
$e_3$ direction. The aim of the section is to present some
corresponding $\Gamma$-convergence analysis. We discuss the micromagnetic and the magnetoelastic
limit separately.
\subsection{ Micromagnetic limit}
We shall prove the convergence of minimizers
of $\bar E^{\rm mag}_h$
to minimizers of some limiting
energy $E^{\rm mag}_0$ as $h\to 0$. Our argument corresponds to an
extension of the analysis by {\sc Gioia \& James}
\cite{goia-james}, who investigated the case of a thickness- and space-independent
magnetic potential $\bar \varphi$ in absence of external field,
i.e. $\bar H=0$. We shall set the result
within the classical
$\Gamma$-convergence frame \cite{braides,dalmaso}. Considering the
standard rescaling with $Z_h:={\rm diag}(1,1,1/h)$ and the mapping
$x\mapsto Z_hx$, we associate to $\bar m:\O_h\to{\mathbb R}^3$ a magnetization
$m:\O:=\O_1\to{\mathbb R}^3$, to $\bar \xi:{\mathbb R}^3\to {\mathbb R}$ the
rescaled magnetostatic potential $\xi:{\mathbb R}^3\to {\mathbb R}$,
to $\bar H:\O_h\to{\mathbb R}^3$ the external field $ H:\O\to{\mathbb R}^3$,
and to $\bar \varphi_h: \Omega_h\times m_{\textrm{sat}} S^2\to [0,\infty) $ the
rescaled magnetic potential $\varphi_h: \Omega\times m_{\textrm{sat}} S^2 \to [0,\infty) $
defined as
\begin{align*}
&m(Z_hx):=\bar m(x), \ \ H(Z_hx):=\bar H(x), \ \ \varphi_h(Z_hx,m):=
\bar \varphi_h(x,m), \\
&\qquad \mbox{ and} \ \ \xi(Z_hy):=\bar \xi(y)\ \ \forall
x\in\O_h, \ y\in{\mathbb R}^3, \ m \in m_{\textrm{sat}} S^2.
\end{align*}\color{black}
Correspondingly, we define the set $\mathcal M : = \mathcal M_1$. By using the summation convention we can express
$$
\nabla\bar m(x)=m,_{i}(Z_hx)\otimes e_i+ \frac1h m,_{3}(Z_hx)\otimes e_{3}\ \text{ where }i=1,2.$$
Moreover, we define the {\it planar} components of the magnetization and of the gradients as
\begin{align*}
&m_p:=m_ie_i=(m_1,m_2), \ \ \nabla_p m:=m,_i\otimes e_i=(m,_1,m,_2), \\
&\nabla_p \xi:= \xi,_ie_i=(\xi,_1,\xi,_2).
\end{align*}
By exploiting this rescaling and notation, we can
equivalently write the energy $\bar E^{\rm mag}_h(\bar m)$ in terms of $m$ as $\bar E^{\rm mag}_h(\bar m)=E^{\rm mag}_h(m)$ where
\begin{align*}
&E^{\rm mag}_h(m):= \frac{\alpha}{|S|} \int_\O\left(|\nabla_p m(z)|^2 +
\frac{1}{h^2}|m,_{3}(z)|^2\right)\, {\rm d} z \\
&+ \frac{1}{|S|} \int_\O\left(\varphi_h(z, m(z))-H(z)\cdot m(z) +\frac{1}{2}\left(\nabla_p \xi(z)\cdot m_p(z)+\frac1h \xi,_3(z)m_3(z)\right)\right)\,{\rm d} z\
\end{align*}
where the relationship between the magnetization and the stray field
is given by the Maxwell equation in the whole space
\begin{align}\label{stray-scaled}
\nabla_p\cdot(-\nabla_p \xi + m_p\chi_\O)+\frac{1}{h}\frac{\partial}{\partial z_3}\left(-\frac1h \xi,_3+m_3\chi_\O\right)=0 .
\end{align}
Moreover, $m$ will be required to satisfy the saturation
constraint $|m|=m_{\textrm{sat}}$ a.e.~in $\O$.
The following result can be found in \cite[Prop.~4.1]{goia-james}.
\begin{lemma}\label{gioia}
Let $\widehat m_h\chi_\O\to \widetilde m\chi_\O$ in $L^2({\mathbb R}^3;{\mathbb R}^3)$ as $h\to
0$, let $|\widehat m_h|=m_{\textrm{sat}}$ a.e.~in $\O$ and let $\widehat\xi_h$ be the
solution to \eqref{stray-scaled} corresponding to $\widetilde m_h\chi_\O$.
Then, we have that $\|\nabla\widehat\xi_h\|_{L^2({\mathbb R}^3;{\mathbb R}^3)}\to 0$ and $\|h^{-1}\widehat\xi_{h,3}-\widetilde m,_3\|_{L^2({\mathbb R}^3)}\to 0$.
Moreover,
\begin{equation}
\lim_{h\to 0} \frac12 \int_\O\left(\nabla_p \widehat\xi_h(z)\cdot (\widehat
m_{h})_p(z)+\frac1h \widehat\xi_{h,3}(z)\,
\widehat m_{h3}(z)
\right)\,{\rm d}
z= \frac12 \int_\O(\widetilde m_3(z))^2\,{\rm d} z .\label{eq_lemma}
\end{equation}
\end{lemma}
The next result describes the limiting micromagnetic energy. It can be
basically
found in \cite[Thm.~4.1]{goia-james} although for $H=0$. The
extension to $H\ne 0$ is straightforward since the
Zeeman term is linear in the magnetization. We shall use
the notation $\widehat H(z_1,z_2):= \int_0^1 H(z_1,z_2,s){\rm d} s$
for $(z_1,z_2)\in S$ (note however that in most applications the
external field can be considered to be constant in~$\O$). With
respect to \cite{goia-james}, we present here a rephrasing of the
result in terms of $\Gamma$-convergence of the micromagnetic energies
\cite{braides,dalmaso}. In particular, we
need to be postulating some limiting behavior of the sequence
$\varphi_h$. Instead of heading to maximal generality,
we prefer to present here a specific yet relevant case by assuming the decomposition
\begin{equation}
\varphi_h(z,m) = f(h)\varphi_p(z_p,m_p) +
\varphi_3(z,m)\label{ansatz}
\end{equation}
where $\varphi_p: S \times \{m \in {\mathbb R}^2 \ | \ |m| \leq m_{\textrm{sat}}\} \to [0,\infty)$ and $\varphi_3:
\Omega \times m_{\textrm{sat}} S^2 \to [0,\infty)$ are continuous and $f: (0,1) \to [0,\infty)$ decreases.
Under suitable coercivity assumptions on $\varphi_p$, the first term
in the above right-hand side penalizes the planar components of
$m$. This would ideally correspond to the observed behavior of some
ultrathin films showing a very strong magnetic anisotropy in the off-plane
direction \cite{Bruno88,Schulz94}.
Along with the above structural {\it Ansatz} \eqref{ansatz}, the magnetic potential
$\varphi_h$ converges pointwise and monotonically to
$$\varphi_0(z,m) = f(0+)\varphi_p(z_p,m_p) + \varphi_3(z,m)$$
where the term $f(0+)\varphi_p$ has to be intended as the
constraint $\{\varphi_p=0\}$ in case $f(0+) = \infty$. Along with this
convention, we have
the $\Gamma$-convergence of the magnetic energy terms in terms of
the strong topology of $L^2$.
\begin{proposition}[$\Gamma$-convergence of the micromagnetic energies]\label{gamma1}
$E^{\rm mag}_h$ $\Gamma$-converges strongly in $L^2$ to $E^{\rm
mag}_0$ given by
$$
E^{\rm mag}_0(m):=
\begin{cases}
E^{\rm mag}_p(m)&:=\displaystyle\frac{1}{|S|}
\int_S\left(\alpha| \nabla_p m|^2+ \varphi_0(z, \color{black} m)- \widehat H\cdot
m+\frac12m_3^2\right)\,{\rm d} z_1{\rm d} z_2 \\
&\hspace{8mm}\mbox{if $m\in W^{1,2}(\O;{\mathbb R}^3)$, $|m|=m_{\rm sat}$, and $m,_3=0$},\\
+\infty &\mbox{otherwise.}
\end{cases}
$$
\end{proposition}
\begin{proof}
The existence of a recovery sequence follows by pointwise
convergence. Let $m \in \mathcal M$ with $m,_3=0$. Then, the
constant sequence $m_h = m$ satisfies $$\lim_{h\to 0}E^{\rm
mag}_h(m)= E^{\rm mag}_0(m)$$ due to Lemma \ref{gioia}. If on the
contrary $m,_3\not =0$ then $E^{\rm
mag}_h(m) \to \infty$.
Let now $m_h \in \mathcal M$ converge strongly in $L^2$ to $m\in \mathcal M$.
As we are interested in checking that $\liminf_h E^{\rm mag}_h(m_h)
\geq E^{\rm mag}_0(m)$ we may assume with no loss of generality that
$\liminf_h E^{\rm mag}_h(m_h)<\infty$ or even that $ E^{\rm
mag}_h(m_h)$ is uniformly bounded. This entails in particular that
$m,_3=0$. It is hence sufficient to use \eqref{eq_lemma} in order
to get the liminf inequality.
\end{proof}
\subsection{ Magnetoelastic limit }
Let us here consider the thin-film limit for the magnetoelastic problem. We shall rescale the magnetoelastic energy
$\bar E^{\rm elas}_h(\bar u,\bar m)$
and obtain a magnetoelastic Kirchhoff-Love
plate theory. In particular, for $x\in\Omega_h$ we let
$$\bar u(x)=:Z_hu(Z_hx)= Z_hu(x_1,x_2,x_3/h)$$
so that $u:\Omega \to {\mathbb R}^3$. Correspondingly, we define the set $\mathcal U : = \mathcal U_1$.
It follows that, for $i,j=1,2,$
$$
\varepsilon_h(u):=Z_h\varepsilon (\bar u)Z_h= \left(\begin{array}{ccc}
\varepsilon(u)_{ij} & \displaystyle\frac{1}{h}\varepsilon(u)_{i3}\\[3mm]
\displaystyle\frac{1}{h}\varepsilon(u)_{i3} &\displaystyle\frac{1}{h^2}\varepsilon(u)_{33} \\
\end{array}\right).
$$
Analogously the scaling for $\varepsilon^{\rm mag}$ will be
$$\varepsilon^{\rm mag}_h(m):=Z_h \varepsilon^{\rm mag}(\bar m(Z_h))Z_h,$$
so that, for $i,j = 1,2$,
$$
\varepsilon^{\rm mag}_h(m):=\left(\begin{array}{ccc}
(\varepsilon^{\rm mag}(m))_{ij} & \displaystyle\frac{1}{h}(\varepsilon^{\rm mag}( m))_{i3}\\[3mm]
\displaystyle\frac{1}{h}(\varepsilon^{\rm mag}(m))_{i3} & \displaystyle\frac{1}{h^2}(\varepsilon^{\rm mag}(m))_{33} \\
\end{array}\right).
$$
As to boundary conditions, we consider
\begin{align}\label{K-L-boundary}
u^{\rm Dir}\in \mathcal K:=\{u \in W^{1,2}(\O;{\mathbb R}^3)\ : \ \varepsilon(u)_{3i}=\varepsilon(u)_{i3}=0 \mbox{ for } i=1,2,3 \}\end{align}
and set (analogously to the choice for $u$)
$\bar u^{\rm Dir}_h(x):=Z_h u^{\rm Dir} (Z_hx)$ for $x\in\O_h$.
The space $\mathcal K$ in \eqref{K-L-boundary} represents the
admissible displacements for Kirchhoff-Love plates. In particular, $u \in \mathcal K$ entails
$$ u_{1,3}+u_{3,1} = u_{2,3}+u_{3,2}=u_{3,3}=0.$$
Namely $u_3$ is constant in direction $e_3$ and $u_1, u_2$ are affine
in direction $e_3$.
We shall define the energy $E^{\rm elas}_h$ on the rescaled
domain $\O$
via $E^{\rm elas}_h(u, m)=\bar E^{\rm elas}_h(\bar u,\bar m)$. In
particular, we have
\begin{align}
E^{\rm elas}_h(u, m)
:=\frac{1}{2|S|}\int_{\O}\mathbb{C}\big(\varepsilon_h(u{+}u^{\rm Dir}){-}\varepsilon_h^{\rm mag}(m)\big){:}\big(\varepsilon_h(u{+}u^{\rm Dir}){-}\varepsilon_h^{\rm mag}(m)\big)\,{\rm d} z .
\end{align}
Let us denote the set of admissible states by
\begin{align}
\mathcal{Q}:= \left\{(u,m)\in\mathcal U \times\mathcal M \ : \
\varepsilon(u)_{i3}=\varepsilon^{\rm mag}(m)_{i3} \ \text{for} \ i=1,2,3\right\}.\nonumber
\end{align}
Then, $E^{\rm elas}_h $ admits a minimizer in
$\mathcal{Q}$. Moreover, the component $u$ of such minimizer depends
linearly on the component $m$. In particular, given $m$, the
displacement $u$ is uniquely
determined.
In order to discuss the limiting case $h\to 0$, by following
\cite{braides} or \cite{freddi-paroni-zanini} we define, for
$i,j,k,\ell=1,2$, the limiting elasticity tensor $\mathbb{C}^0$
as
\begin{align}
\mathbb{C}^0_{ijk\ell}:=\mathbb{C}_{ijk\ell}-\frac{\mathbb{C}_{ij33}\mathbb{C}_{k\ell33}}{\mathbb{C}_{3333}} .
\end{align}
For all $ A\in{\mathbb R}^{2\times 2}$ we let the quadratic form $Q:{\mathbb R}^{2\times 2} \to [0,\infty)$ be defined as
$$Q(A):=\min_{a\in{\mathbb R}^{2\times 1 },b\in{\mathbb R}}\mathbb{C}\left(\begin{array}{ccc}
A & a\\
a^\top & b \\
\end{array}\right): \left(\begin{array}{ccc}
A & a\\
a^\top & b \\
\end{array}\right) .
$$
One readily checks that the minimum is achieved at
$$a=0 \ \ \text{and} \ \ b= - \frac{\mathbb{C}_{ij33} A_{ij}}{\mathbb{C}_{3333}} .$$
In particular, we have that
$Q(A)=\mathbb{C}^0 A{:} A$ for all $A\in {\mathbb R}^{2\times 2}$ and that $Q$ is
uniformly convex on ${\mathbb R}^{2\times 2}$. Let us use the notation $\varepsilon_p \in {\mathbb R}^{2\times 2}$ in order to indicate the {\it
planar} block of the matrix $\varepsilon \in {\mathbb R}^{3\times 3}$, namely
$(\varepsilon_p)_{ij} = (\varepsilon)_{ij}$ for $i,j
=1,2$. We have the following.
\begin{proposition}[$\Gamma$-convergence of the magnetoelastic energies]\label{Prop-recovery}
For all $u^{\rm Dir} \in \mathcal K$ we have that $E^{\rm elas}_h$ $\Gamma$-converges to $ {E}^{\rm elast}_0$ with
respect to the weak topology of $W^{1,2}$ where
$$
{E}^{\rm elast}_0(u,m):=
\begin{cases}
E^{\rm elas}_p(u,m) &:= \displaystyle\frac{1}{2|S|} \int_{
\O }Q\big(\varepsilon_p(u){+}\varepsilon_p(u^{\rm Dir}(t)){-}\varepsilon_p^{\rm
mag}(m)\big) {\rm d} z\\
\qquad & \quad \mbox {if $(u,m)\in \mathcal{Q}$},\\
+\infty &\quad \mbox{otherwise.}
\end{cases}
$$
\end{proposition}
\begin{proof}
Let $(u_h,m_h)\to (u,m)$ weakly in $W^{1,2}$ and assume with no
loss of generality that $\liminf_{h\to 0}
E^{\rm elas}_h(u_h,m_h)<\infty$ or even, possibly extracting but not
relabelling, that $E^{\rm elas}_h(u_h,m_h)$ are uniformly bounded. Then, since
$u^{\rm Dir} \in \mathcal K$, the
limit $(u,m)$ belongs necessarily to $\mathcal{Q}$. Hence, the
definition of $Q$ and its lower semicontinuity imply
that $\liminf_{h\to 0}
E^{\rm elas}_h(u_h,m_h)\ge {E}^{\rm elas}_p(u,m)$.
On the other hand, consider $(u,m)\in\mathcal{Q}$. Define $b\in
L^2(\O)$ via
\begin{align}\label{b}
b:= - \frac{\mathbb{C}_{ij33} (\varepsilon(u{+}u^{\rm Dir})_{ij}-\varepsilon^{\rm mag}(m)_{ij})}{\mathbb{C}_{3333}}\ \end{align}
and a sequence $\{\psi_h\}_{h>0}\subset C^\infty_0(\O)$ such that
$\psi_h\to b$ in $L^2(\O)$ and $h\nabla\psi_h\to 0$ in
$L^2(\O;{\mathbb R}^3)$. Let $\phi_h$ be such that
$\phi_{h,3} := \psi_h$ in $\Omega$.
Define for $h>0$ and i=1,2,3,
$$
\widetilde u_{hi} :=
\begin{cases}
u_i & \mbox{ for $i =1,2$,}\\
u_i +h^2\phi_h & \mbox{ for $i=3$.}
\end{cases}
$$
Hence,
\begin{align}
&\varepsilon_h(\widetilde u_h{+}u^{\rm Dir})-\varepsilon_h^{\rm
mag}(m)\nonumber\\
&= \left(\begin{array}{cc}
\varepsilon(\widetilde u_h{+}u^{\rm Dir})_{p}{-}\varepsilon^{\rm
mag}(m)_{p} & \displaystyle\frac{(\varepsilon(\widetilde u_h{+}u^{\rm Dir})_{i3}{-}\varepsilon^{\rm mag}(m)_{i3})}{h}{+} \frac{h}{2}\frac{\partial\phi_h}{\partial x_i}\\
\displaystyle\frac{(\varepsilon(\widetilde u_h{+}u^{\rm
Dir})_{3i}{-}\varepsilon^{\rm mag}(m)_{3i})}{h}{+}\displaystyle \frac{h}{2}\frac{\partial\phi_h}{\partial x_i}&\!\!\!\! \psi_h\\
\end{array}\right).\nonumber
\end{align}
As $u^{\rm Dir} \in \mathcal K$, we have
$\varepsilon(u{+}u^{\rm Dir})_{i3}-\varepsilon^{\rm mag}(m)_{i3}=0$. In
particular, the terms with factor $1/h$ vanish. Then,
passing to the $\liminf$ for $h \to 0$ we readily get that $E^{\rm
elas}_h(\widetilde u_h, m) \to E^{\rm elas}_p(u,m)$. In particular,
$(\widetilde u_h, m) $ is a recovery sequence for $(u,m)$.
In case $(u,m) \not \in \mathcal{Q}$, the same choice $(\widetilde u_h, m)
$ entails $E^{\rm elas}_h(\widetilde u_h, m)\to \infty$. Namely,
$(\widetilde u_h, m)$ still provides a recovery sequence.
\end{proof}
The rescaled total magnetoelastic energy $E_h$ is defined
on the domain $\O$ via $E_h(u,m) = \bar E_h(\bar u, \bar m)$. In particular, we have
$$
E_h(u,m) = E^{\rm mag}_h(m) + E^{\rm elas}_h(u,m).
$$
\section{ Quasistatic evolution}\label{evo}
Let us now turn to the analysis of the quasistatic
evolution case. The aim here is to introduce and analyze a rate-independent
model for a magnetoelastic Kirchhoff-Love plate. We obtain this by
dimension reduction, by passing to the limit in $h>0$ for a
three-dimensional magnetoelastic evolution model. In particular, we
detail in Subsection \ref{analysis} the quasistatic evolution problem
for the three-dimensional specimen and discuss in Subsection \ref{KL}
the evolutive thin-film limit constituing the plate model.
\subsection{ Quasistatic evolution in the bulk}\label{analysis}
\setcounter{equation}{0}
By possibly assuming the Dirichlet datum $\bar u^{\rm Dir}_h$ and/or
the external field $H$ to change with time, the minimizers $(\bar
u,\bar m)$ of $\bar E_h$ evolve as well. In order to prescribe a
suitable evolution law, we postulate magnetic dissipation. In
particular, by assuming that the changes in the data are so slow that
inertial effects can be neglected, we assume that the (time-dependent)
state of the system $t \mapsto (\bar u(t),\bar m(t))$ solves relations \eqref{sys} on
$[0,T]$ (and in a
suitable variational sense, see below) where however the static
relation \eqref{m} is replaced by the rate-independent inclusion
$$
\partial \psi(\bar m_t) - \alpha \Delta \bar m + \nabla
_{\bar m}\bar \varphi_h(x, \bar m) +\frac12 \nabla
\bar
\xi \ni \bar H \quad \text{in} \ \Omega_h \times (0,T).$$
The symbol $\partial$ above indicates the subdifferential in the sense
of convex analysis and $\psi(\bar m_t)$ measures the infinitesimal
dissipation involved in the process. As the thickness $h$
decreases, an additional magnetic anisotropy effect arises. While bulk materials $h=1$
show isotropic dissipation, in the thin-film limit $h \to 0$
anisotropic dissipation can be observed
\cite{uno,due,tre}. In particular, in some regimes the
dissipation tends to be larger for processes involving off-planar
magnetizations. We shall take this into account by choosing
$$\psi(\bar m_t) = R_p |\bar m_{p,t}| + R_3(h)|\bar m_{3,t}|.$$
Here, $R_p >0$ is an energetic yield
limit for evolution in the plane \cite{hubert-schaefer} which we
assume to be independent of the film thickness, for simplicity. On
the other hand, the function $h \mapsto R_3(h) >0$ models
anisotropic effects in
the $e_3$ direction which are
observed to be thickness-dependent \cite{quattro}. \color{black} We shall here limit ourselves in assuming that the right limit $R_3(0_+)$ exists and is
finite. Note nonetheless that the case
$R_3(0_+)=\infty$, imposing indeed $\bar m_{3,t}=0$, could be considered as well.
The latter
equation corresponds to the postulate that the
energy released by changing the state of the system
from $(\bar u^1,\bar m^1)$ to $(\bar u^2,\bar m^2)$ is given by the
simple form
\begin{align}
\bar D_h(\bar m^1,\bar m^2):= \frac{1}{|S|h}\int_{\O_h} \left(R_p|\bar
m^1_p{-}\bar m^2_p| + R_3(h)|\bar
m^1_3{-}\bar m^2_3|\right){\rm d} x. \nonumber
\end{align}
Note that the
dissipation $\bar D_h$ is positively $1$-homogeneous and, correspondingly, the evolution will be
rate-independent. In particular, energy will be dissipated by purely
hysteretic losses.
For the sake of later convenience, let us reformulate the problem in
the fixed reference configuration $\Omega$. This amounts in
considering the energies $ E_h^{\rm mag}$, $ E_h^{\rm elas}$, and $
E_h $ (here assumed to be depending on time as well, without
introducing new notation) and the dissipation
$D_h(m^1,m^2)=\bar D_h(\bar m^1,\bar m^2)$ so that
$$
D_h( m^1, m^2)=\frac{1}{|S|}\int_{\O} \left(R_p|
m^1_p{-}m^2_p| + R_3(h)|
m^1_3{-}m^2_3|\right){\rm d} x.\color{black}
$$
Assume to be given time-dependent boundary datum $t \in [0,T]\mapsto \bar
u^{\rm Dir}(t) \in W^{1,2}(\O,{\mathbb R}^3)$ and
external field $t \in [0,T]\mapsto H(t) \in L^1(\O_h;{\mathbb R}^3)$. We are
interested in proving the existence of a quasistatic evolution $t\in [0,T] \mapsto ( u_h(t),m_h(t)) \in
\mathcal{Q} $
in the form of the
so-called {\em energetic formulation} \cite{mielketheil}. Given
some suitable initial datum
$(u^0, m^0) \in\mathcal{Q}$
we define it as follows.
\begin{definition}[Energetic solution in the bulk]\label{Def:ES-bulk}
An \emph{energetic solution} of the quasistatic evolution in
the bulk
is a trajectory $t\in[0,T]\mapsto ( u_h(t), m_h(t))
\in \mathcal{Q} $ such that $( u_h(0), m_h(0))=( u^0, m^0)$ and, for every $t\in [0,T]$,
\begin{align}
& E_h(t, u_h(t), m_h(t))\leq E_h(t,\widehat u,\widehat m)+
D_h( m_h(t),\widehat m)\quad \forall\, (\widehat u,\widehat m)\in \mathcal{Q} \tag{S}\label{S}\\[2mm]
& E_h(t, u_h(t), m_h(t))+\hbox{\rm Diss}_{ D_h}( m_h,[0,t])
\nonumber\\
&\qquad = E_h(0, u^0, m^0) +\int_0^t \partial_t E_h(s, u_h(s), m_h(s)){\rm d}s\tag{E}\label{E}
\end{align}
where $\hbox{\rm Diss}_{ D_h}( m_h,[0,t])$ is the \emph{total dissipation} on $[0,t]$ defined by
\begin{equation}\label{dissi}\hbox{\rm Diss}_{ D_h}( m_h,[0,t]):=\sup\left\{\sum_{i=1}^{N} D_h( m_h(t^i), m_h(t^{i-1})) \right\},
\end{equation}
the supremum being taken over all partitions $ \{0=t^0 <t^1<
\ldots<t^N=t\}$ of $[0,t]$.
\end{definition}
The two conditions \eqref{S}-\eqref{E} in the definition of
energetic solution have an immediate
mechanical interpretation. Condition \eqref{S} is a {\it global
stability} criterion: Transitions from the actual state $(
u(t), m(t))$ to some possible competitor state $(\widehat u,\widehat m)$ is not
energetically favored in the sense that the energy gain is
compensated by the dissipation cost.
For later notational convenience, we define the set of {\it stable states} at time $t\in [0,T]$ as
\begin{align*}
\mathcal{S}_h(t)&:=\Big\{( u_h, m_h)\in
\mathcal{Q}: \
E_h(t, u_h, m_h)\leq E_h(t,\widehat u,\widehat m){+} D_h(
m_h,\widehat m),\ \forall\,(\widehat u,\widehat m)\in \mathcal{Q}
\Big\}
\end{align*}
so that condition \eqref{S} equivalently reads $(
u_h(t), m_h(t))\in \mathcal{S}_h(t)$ for all $t \in [0,T]$. The scalar
equation \eqref{E} is nothing but {\it energy conservation}: It
expresses the balance between current and dissipated energy
(left-hand side)
and initial energy plus work of external actions
(right-hand side).
Let us close this section by recording an existence result
for quasistatic evolutions in three dimensions.
\begin{theorem}[Existence for the quasistatic evolution in
the bulk]\label{thcv2} Let $h>0$. Assume to be given $H\in C^1([0,T]; L^1(\O;{\mathbb R}^3))$, $ u^{\rm
Dir}_h\in C^1([0,T]; W^{1,2}(\O;{\mathbb R}^3))$, and
$( u^0, m^0)\in\mathcal{S}_h(0)$. Then, there exists an energetic
solution $( u_h, m_h)$ for the quasistatic evolution problem.
\end{theorem}
We shall not report here a proof of Theorem \ref{thcv2} as it may be
readily obtained in the frame of the by now classical existence theory
for energetic solutions by {\sc Mielke \& Theil}
\cite{mielke01,mielketheil}. Indeed, it is sufficient to point out
that $E_h$
has bounded (hence weakly compact) sublevels in $\mathcal{Q}$,
that $D_h$ is continuous with respect to the same topology, and that
the power $\partial_t E_h$ is well behaved in
order to apply, for instance, \cite[Thm. 5.2]{mielke01}.
\subsection{ Quasistatic evolution of the magnetoelastic thin-film}\label{KL}
Let us now come to the description of the magnetoelastic thin-film,
which results in a
Kirchhoff-Love plate model. We shall derive this by taking the limit
$h\to 0$ in the three-dimensional evolution model. The state of the material will be
described by the pair
\begin{align*}
(u,m)&\in \haz{\mathcal{Q}} := \{ (u,m) \in \mathcal{Q} \ \
\text{such that} \ m_{{,3}}=0\}\\
& = \{(u,m) \in W^{1,2}_0(\O,{\mathbb R}^3)\times W^{1,2} (\O,{\mathbb R}^3) \
\text{such that }\\
&\hspace{5mm}\varepsilon_{i3}(u) = \varepsilon^{\rm mag}(m)_{i3} \ \text{for} \ i=1,2,3, \ |m|=m_{\textrm{sat}}, \ m,_3=0 \
\text{a.e.}
\}
\end{align*}
and its statics will
correspond to the minimization of the thin-limit energies of Section
\ref{dimension-reduction}.
In the following, the boundary datum and the external
field are time-dependent and will be driving the quasistatic
evolution of the medium. Correspondingly, we will indicate
time-dependence in the total energy of the medium as
$ E_0(t,u,m) = E^{\rm mag}_0(t,m) + E^{\rm elas}_0 (t,u,m)$. Note that
$E_0(t,\cdot)$ is finite on $\haz{\mathcal{Q}}$.
As for the dissipation,
for all $m^1,m^2\in\mathcal{M}$ which are hence constant in
the direction $e_3$, we define
\begin{align}\label{dissipation}
D_0(m^1,m^2):= \frac{1}{|S|} \int_S \left(R_
|m^1_p(z){-}m^2_p(z)| + R_3(0_+)|m^1_3{-}m^2_3|\right){\rm d} z . \end{align}
Owing to these definitions, the quasistatic evolution problem
for the magnetoelastic thin film can be reformulated in
terms of a rate-independent evolution driven by the potentials
$( E_0,D_0)$. As before, we shall be interested in energetic
solutions.
\begin{definition}[Energetic solution for the thin film]\label{Def:ES-bulk2}
An \emph{energetic solution} of the quasistatic evolution for
the magnetoelastic thin film
is a trajectory $t\in[0,T]\mapsto ( u(t), m(t)) \in
\haz{\mathcal{Q}} $ such that $( u(0), m(0))=( u^0, m^0)$ and, for every $t\in [0,T]$,
\begin{align}
& E_0(t, u(t), m(t))\leq E_0(t,\widehat u,\widehat m)+ D_0( m(t),\widehat m)\quad
\forall\, (\widehat u,\widehat m)\in \haz{\mathcal{Q}} \tag{S2}\label{S2}\\[2mm]
& E_0(t, u(t), m(t))+\hbox{\rm Diss}_{ D_0}( m,[0,t])\nonumber\\
&\qquad
= E_0(0, u^0, m^0) +\int_0^t \partial_t E_0(s, u(s), m(s)){\rm d}s\tag{E2}\label{E2}
\end{align}
where $\hbox{\rm Diss}_{ D_0}( m,[0,t])$ is the \emph{total
dissipation} on $[0,t]$ defined analogously to
$\hbox{\rm Diss}_{ D_h}$, but starting from the
dissipation $D_0$.
\end{definition}
Let us denote by $\mathcal{S}_0(t)$
the set of
stable states at time $t$, namely of
pairs $(u,m) \in \haz{\mathcal{Q}}$ fulfilling \eqref{S2}.
We shall now prove that energetic solutions to the quasistatic
evolution problem for the
magnetoelastic Kirchhof-Love plate exist. Indeed, our result is
stronger, as we prove that sequences of solution of the bulk model
admit subsequences which converge to energetic solution of the
thin-film model. In particular, we provide an approximation
result based on dimension reduction.
\begin{theorem}[Convergence to the thin film]\label{thm}
Let $u^{\rm Dir}\in C^1([0,T];W^{1,2}(\O;{\mathbb R}^3)$, $H\in
C^1([0,T]; L^1({\mathbb R}^3))$, $ (u^0,m^0)\in \mathcal{S}_0(0)$, $ E_h(0,u^0,m^0)\to E_0(0,u^0,m^0)$, and
$\{( u_h, m_h)\}_{h>0}\subset\mathcal{Q}_h$ be a sequence of
energetic solutions of the quasistatic evolution in three dimensions, i.e. solving \eqref{S}-\eqref{E}. Then, for some not relabeled
subsequence we have that $( u_h, m_h)\rightharpoonup (u,m)$ in
$W^{1,2}(\O;{\mathbb R}^3)\times W^{1,2}(\O;{\mathbb R}^3)$ where $(u,m)$ is an
energetic solution for the plate, i.e. solving\eqref{S2}-\eqref{E2}.
\end{theorem}
In order to show that an energetic solution of the bulk material converges
to an energetic solution to a plate we apply the abstract
strategy introduced in
\cite{mrs}. We shall not provide here a detailed proof, but
rather comment on two crucial points of the argument. The first of
these points concerns functional convergence. In particular, we shall
establish a specific {\it evolutive
$\Gamma$-convergence} notion, adapted to rate-independent
evolutions. Indeed, the theory relies on the verification of
{\it two separate} $\mathop{\Gamma\text{--}\mathrm{liminf}}$ inequalities
\begin{equation}
\label{gliminf}
E_0 \leq \mathop{\Gamma\text{--}\mathrm{liminf}}_{h \to 0} E_h, \qquad D_0 \leq \mathop{\Gamma\text{--}\mathrm{liminf}}_{h
\to 0} D_h
\end{equation}
as well as on a {\it mutual recovery sequence} condition. The
$\mathop{\Gamma\text{--}\mathrm{liminf}}$ inequality for $ E_h$ follows by easily adapting the results of Section
\ref{dimension-reduction} to the present time-dependent case. On
the other hand, the $\mathop{\Gamma\text{--}\mathrm{liminf}}$ inequality for $D_h$ is immediate as
$R_3(h) \to R_3(0_+) \geq 0$.
The
following lemma entails the existence of a mutual recovery sequence.
\begin{lemma}[Mutual recovery sequence]\label{joint}
Let $(t_h,u_h,m_h) \rightharpoonup (t,u,m)$ in $[0,T]\times
\mathcal{Q}$, and $(\widehat u,\widehat m)\in\haz{\mathcal{Q}}$. Then, there exist $(\widehat u_h,\widehat m_h)
\rightharpoonup (\widehat u,\widehat m) $ such that
\begin{align}
\limsup_{h\to 0} ( E_h(t_h, \widehat u_h, \widehat m_h) +D_h(m_h,\widehat
m_h))\le E_0(t,\widehat u,\widehat m)+D_0(m,\widehat m).
\end{align}
\end{lemma}
\begin{proof}
For all $h>0$, we choose $\widehat
m_h:=\widehat m$ and $\widehat u_h$ as in the proof of
Proposition~\ref{Prop-recovery}. The claim then follows by the
continuous convergence of $D_h$ to $D_0$
with respect to the strong $L^2$-convergence of its
arguments.
\end{proof}
A second crucial point for the possible application of the
abstract argument of \cite{mrs} consists in the convergence proof of
the power of the energy functionals. We shall argue here
in the same spirit of \cite{freddi-paroni-zanini}.
\begin{lemma}\label{powers}
Let $u^{\rm Dir}\in C^1([0,T];W^{1,2}(\O;{\mathbb R}^3))$ and $H\in
C^1([0,T];L^1({\mathbb R}^3))$.
Let $(t,u,m)\in(0,T)\times\mathcal Q$ and assume that there is a
sequence $(t_h,u_h,m_h)\in (0,T)\times \mathcal Q$ such that
$(u_h,m_h)\in\mathcal{S}_h(t_h)$ and $t_h\to t$, $u_h\rightharpoonup
u$ and $m_h\rightharpoonup m$ in $W^{1,2}(\O;{\mathbb R}^3)\times
W^{1,2}(\O;{\mathbb R}^3)$. Then, we have the convergence of the energies
and the corresponding powers
\begin{align*}
&E_h(t_h,u_h,m_h)\to E_0(t,u,m),\\
&\partial_t E_h(t_h,u_h,m_h)\to\partial_t E_0(t,u,m).
\end{align*}
\end{lemma}
\begin{proof}
Let us first show the convergence of the energies. The inequality $$\liminf_{h\to
0} E_h(t_h,u_h,m_h)\ge E_0(t,u,m)$$ follows from the
already checked $\Gamma$-convergence of the functionals,
i.e. Propositions~\ref{gamma1} and \ref{Prop-recovery}. To check for the other inequality we use Lemma~\ref{joint} and stability.
Indeed, by choosing $\widehat u:=u$, $\widehat m:=m$ and
letting $(\widehat u_h, \widehat m_h)$ be the corresponding sequences from Lemma~\ref{joint} we have
\begin{align*}
\limsup_{h\to 0} E_h(t_h,u_h,m_h)&\le \limsup_{h\to 0}( E_h(t_h, \widehat u_h,
\widehat m_h) +D_h(m_h,\widehat m_h)) \\
&\le E_0(t,\widehat u,\widehat m) + D_0(m,m)= E_0(t, u, m),
\end{align*}
where the first inequality follows from the stability $(u_h,m_h)
\in \mathcal{S}_h(t_h)$.
Let us now compute the power of $ E_h$ as
\begin{align*}
\partial_t
E_h(t_h,u_h,m_h)&=\int_{\Omega_h}\mathbb{C}(\varepsilon(u_h){+}\varepsilon(u^{\rm
Dir}(t_h)){-}\varepsilon^{\rm mag}(m_h)):\varepsilon(\dot u^{\rm
Dir}(t_h))\, {\rm d} x\\
&-\int_{\Omega_h} \dot H(t_h)\cdot m_h\,{\rm d} x.
\end{align*}
An analogous expression holds for $\partial_t E_0(t,u,m)$. The
convergence of the first term in the expression of $\partial_t
E_h(t_h,u_h,m_h)$ can be proved as in \cite{freddi-paroni-zanini}
while the convergence of the second term
is immediate by linearity.
\end{proof}
Given the $\Gamma$-convergence of the functionals (Section~\ref{dimension-reduction}) and the powers (Lemma~\ref{powers}) and the
existence of a mutual recovery sequence (Lemma \ref{joint}), it
suffices to remark that $ E_h$ is coercive with respect to the
weak topology of $W^{1,2}(\O;{\mathbb R}^3)\times W^{1,2}(\O;{\mathbb R}^3)$ in
order to obtain Theorem \ref{thm} by applying the abstract theorem \cite[Thm~3.1]{mrs}.
Before closing this discussion let us explicitly note that the
developed technologies would allow also to deduce additional
dimension reduction results. In particular, by neglecting mechanical
effects, one could consider the possibility of deducing a
rate-independent model for the quasistatic evolution of a thin-film
driven by micromagnetic energy. This would constitute an evolutive
counterpart to the static analysis in \cite{goia-james}.
\section*{Acknowledgments} This work was initiated during a visit of
MK in IMATI CNR Pavia. Its hospitality and support is gratefully
acknowledged. US is partially supported by the CNR-JSPS grant {\it VarEvol}.
|
\section{I.~Introduction}
The standard model (SM)-like Higgs boson with mass around 126 GeV \cite{h} reported at the LHC strongly
favors extension beyond the minimal supersymmetric model (MSSM).
A simple idea is adding a SM singlet $S$ to the MSSM, with superpotential
\begin{eqnarray}{\label{NMSSM}}
W=\lambda~SH_{u}H_{d}+\frac{\kappa}{3}S^{3}.
\end{eqnarray}
Due to the Yukawa coupling $\lambda$ in Eq.(\ref{NMSSM}),
the Higgs boson mass can be naturally uplifted to observed value
for moderate value $\lambda \geq 0.5$ at the electroweak (EW) scale.
This specific extension is referred as
the next-to-minimal supersymmetric model (NMSSM) \cite{0910.1785} for $\lambda\leq 0.7$,
or $\lambda$-supersymmetry ($\lambda$-SUSY) \cite{0607332} for $0.7\leq \lambda\leq 2$.
Either the NMSSM or $\lambda$-SUSY is very attractive from the viewpoint of naturalness argument \cite{1112.2703, 1312.0181},
as the stringent tension on stop masses required by the Higgs mass in the MSSM is dramatically reduced.
The origin of upper bound on $\lambda$ above can be seen from the beta function $\beta_{\lambda}$ for $\lambda$.
Since $\beta_{\lambda}$ is dominated by the top Yukawa coupling $y_t$,
the sign of which is positive,
it imposes an upper bound on the EW value of $\lambda$
in order to be still in the perturbative region at high energy scale.
Given this scale near the grand unification (GUT) scale $\simeq 1.0\times 10^{15}$ GeV,
the critical value was found to be $\sim 0.7$ \cite{9801437} in terms of the one-loop beta function
\footnote{In this paper, we follow the convention in \cite{9709356},
and present our $\beta$ functions in the $\overline{DR}$ scheme.},
\begin{eqnarray}{\label{nmssmbeta}}
\beta_{\lambda}=\frac{\lambda}{16\pi^{2}}\left[4\lambda^{2}+3y^{2}_{t}+2\kappa^{2
}-3g^{2}_{2}-\frac{3}{5}g^{2}_{1}\right],
\end{eqnarray}
where $t\equiv\ln (\mu/\mu_{0})$, $\mu$ being the running renormalization scale and $\mu_{0}\equiv 1$ TeV.
On the other hand,
the observed Higgs mass favors large value of $\lambda$ \cite{1310.0459}.
The intuition for solving the tension between the observed Higgs mass and perturbativity
is obvious in the context of SUSY.
The main observation is that Yukawa couplings in superpotential receive only wave-function induced renormalizations due to the protection of SUSY,
in contrast to non-supersymmetric case \footnote{The phenomenon of Yukawa coupling being asymptotically free has been exposed in $\lambda\phi^{4}$ scalar field theory with dimension lower than four by Wilson and Fisher \cite{Wilson},
and in well-known nonlinear sigma model \cite{Polyakov} together with an attempt to obtain a version of four-dimensional supersymmetry \cite{Ketov}.
Similar phenomenon was also addressed by authors of \cite{9406199} in four-dimensional scalar field theory with nonpolynomial potentials.}.
By following the fact that the beta functions of Yukawa couplings are related to the anomalous dimensions of chiral fields \cite{9709356},
and the anomalous dimensions are proportional to quadratic Yukawa couplings with positive coefficients,
the sign of contribution to $\beta$ function due to Yukawa interactions is thus always positive.
A way to alleviate this is introducing hidden super-confining gauge dynamics \cite{0405267, 0311349},
from which Yukawa couplings are asymptotically free,
and $S$ is a composite other than fundamental state,
with large $\lambda$ at low energy as a result of this asymptotical freedom.
In this context the Higgs doublets can be either fundamental \cite{0405267} or composite states \cite{ 0311349}.
In this paper, we will explore $\lambda$-SUSY that stays perturbative up to the GUT scale in alternative way.
In this framework we will obtain a well defined $\lambda$-SUSY on the realm of perturbative analysis,
and operators such as Higgs doublets and singlet $S$ are all fundamental other than composite states.
This is our main motivation for this study.
The study of well defined $\lambda$-SUSY was initially addressed in Ref.\cite{9801437},
in which it was found that for $\tan\beta$ smaller than $\sim 10$,
$\lambda$ is the first Yukawa coupling running into the non-perturbative region at high energy scale.
It was also understood that either introducing new matter fields at the intermediate scale between the weak and GUT scale or adopting smaller EW value $\kappa(\mu_{0})$ can decrease the evolution rate for $\lambda$.
Given an initial EW value $\kappa(\mu_{0})$,
one can derive the upper bound on $\lambda(\mu_{0})$ or vice versa,
once new matters which appear at the intermediate scale are identified explicitly.
The minimal content of new fields only includes the messenger sector
which communicates the SUSY breaking effect to the visible sector,
namely the $\lambda$-SUSY.
Together with the messengers a spontaneously broken $U(1)_X$ gauge group
will be considered as the new fields.
We consider a class of models different from those studied in the literature \cite{0404251},
in which SM fermions and sfermions except the Higgs doublets and $S$ are all charged under the hidden $U(1)_{X}$.
The abelian gauge coupling $g_{X}$ and the $U(1)_X$-breaking scale $M_{X}$
enter into the parameter space as two new free parameters.
In comparison with traditional NMSSM or $\lambda$-SUSY,
in our model the beta function for top Yukawa coupling $\beta_{t}$ receives
new and negative contribution due to the hidden $U(1)_X$ sector,
the magnitude of which is determined by the hidden gauge coupling $g_{X}$ and scale $M_X$.
As long as $g_{X}$ is large enough but still valid for perturbative analysis,
these new effects decrease the slope of $\lambda$ as function of renormalization scale $\mu$ above scale $M_{X}$,
and therefore lead to larger critical value $\lambda(\mu_{0})$.
The paper is organized as follows.
In section 2, in terms of one-loop renormalization group equations (RGEs) for relevant coupling constants
we estimate the critical values of $\kappa(\mu_{0})$ and $\lambda(\mu_{0})$
due to the hidden $U(1)_X$ effect.
In section 3, we revise the phenomenology for large $\lambda$ but small $\kappa$.
Finally we conclude in section 4.
In appendix A, we show the details of the hidden sector
and briefly review collider constraints on the model parameters.
\section{II.~Perturbative $\lambda$-SUSY}
As mentioned in the introduction,
the beta function for Yukawa couplings of superpotential can be easily estimated by using the non-renormalization of superpotential.
Firstly, in the case without hidden $U(1)_{X}$ gauge group
the one-loop beta functions for Yukawa couplings in $\lambda$-SUSY are given by
\footnote{The value of $\tan\beta$ defined as the ratio $\left<H^{0}_{u}\right>/\left<H^{0}_{d}\right>$
is required to be smaller than $\sim 10$ in light of the 126 Higgs mass in the NMSSM or $\lambda$-SUSY.
In \cite{0810.0989} it has been shown that experimental constraints require $\tan\beta \leq 10$ for $\lambda\geq1.0$.
Thus, it is a good approximation to ignore the bottom Yukawa coupling in compared with top Yukawa coupling.},
\begin{eqnarray}{\label{beta0}}
\beta_{\lambda}&=& \frac{\lambda}{16\pi^{2}} \left[ 4\lambda^{2}+2\kappa^{2}+3y^{2}_{t}-3g_{2}^{2}-\frac{3}{5}g^{2}_{1}\right],\nonumber\\
\beta_{y_{t}}&=& \frac{y_{t}}{16\pi^{2}} \left[ 6y^{2}_{t}+\lambda^{2}-\frac{16}{3}g^{2}_{3}-3g^{2}_{2}-\frac{13}{15}g^{2}_{1}\right] ,\nonumber\\
\beta_{\kappa}&=& \frac{\kappa}{16\pi^{2}} \left[6\lambda^{2}+6\kappa^{2} \right].
\end{eqnarray}
The one-loop beta functions for the SM gauge couplings are the same as the MSSM,
\begin{eqnarray}{\label{gauge}}
\frac{\partial}{\partial t} \alpha^{-1}_{i} &=& -\frac{b_{i}}{2\pi}.
\end{eqnarray}
where $(b_{1}, b_{2}, b_{3})=(33/5, 1, -3)$.
In the case with a hidden $U(1)_{X}$ gauge group,
we study the model in which SM fermions and sfermions all carry a hidden $U(1)_{X}$ charge,
with the spontaneously broken scale $M_{X}\simeq 10$ TeV.
The Higgs doublets, however, are singlets of this $U(1)_{X}$ symmetry.
Anomaly free conditions require three hidden matters $X_{1,2,3}$ with the same $U(1)_{X}$ charge added to the model.
For details about the matter representations, see appendix A.
The modifications to RGEs in Eq.(\ref{beta0}) due to the hidden $U(1)_{X}$ effect are given by,
\begin{eqnarray}{\label{beta}}
\delta \beta_{\lambda}&=&0,\nonumber\\
\delta \beta_{y_{t}}&=&-\frac{y_{t}}{16\pi^{2}}\left(\frac{12}{5}a^{2}\cdot g^{2}_{X}\right), \nonumber\\
\delta \beta_{\kappa}&=&0,
\end{eqnarray}
together with
\begin{eqnarray}{\label{gx}}
\beta_{g_{X}}=\frac{g^{3}_{X}}{16\pi^{2}}\left[\frac{3}{5}a^{2}\cdot\left(\frac{39}{2}n_{g}+3\right)\right],
\end{eqnarray}
where $n_{g}=3$ denotes the number of SM fermion generations.
It is obvious that large EW value $g_{X}(\mu_{0})$ favors small $\mid a\mid$.
The RG running for $g_X$ with different values at $\mu_{0}$ is plotted in Fig.1.
In this figure non-perturbative region lies below the red line.
It shows the critical value $g_{X}(\mu_{0})\leq \{2.0, 1.18, 0.82\}$ for $a=\{\frac{1}{6},\frac{1}{3},\frac{1}{2}\}$, respectively.
Given the upper bound on $g_{X}(\mu_{0})$,
the upper bound on Yukawa coupling $\lambda$ can be estimated
in terms of RGEs in Eq.(\ref{beta}) and Eq.(\ref{beta0}).
The input parameters related to the critical value $\lambda(\mu_{0})$ are composed of
\begin{eqnarray}{\label{input}}
\{y_{t}(\mu_{0}),~\kappa(\mu_{0}),~g_{X}(\mu_{0}),~M,~n_{5\bar{5}}\},
\end{eqnarray}
where $M$ denotes the messenger mass scale,
$n_{5\bar{5}}$ represents the number of $\bar{5}+5$ representation of $SU(5)$ for messengers.
We will use the updated pole mass of top quark $m_{t}=174$ GeV \cite{top}
instead of $m_{t}=180$ GeV in \cite{9801437} for our analysis,
and take the criteria that the theory enters into the non-perturbative region
when any of coupling constants in the theory is bigger than $\sim 4\pi$.
\begin{center}
\begin{figure}
\includegraphics[width=0.45\textwidth]{gx.eps}
\caption{Critical value $g_{X}(\mu_{0})$ for $a=\{\frac{1}{6},\frac{1}{3},\frac{1}{2}\}$, respectively.}
\end{figure}
\end{center}
We want to emphasize that $(i)$, below scale $M_{X}$,
the modifications to RGEs arising from $U(1)_{X}$ can be ignored.
$(ii)$, above scale $M$,
coefficients $b_{i}$ in beta functions for SM gauge couplings should change as $b_{i}\rightarrow b_{i}+n_{5\bar{5}}$.
Fig. 2 shows the modifications to the upper bound on $\kappa(\mu_{0})$ due to the hidden $U(1)_{X}$ sector
given fixed $\lambda(\mu_{0})$.
We choose the initial EW value $\lambda(\mu_{0})=1.0$, $g_{X}(\mu_{0})=0.82$ for $a=1/2$,
and messenger paramaters $M=10^{7}$ GeV and $n_{5\bar{5}}=\{1,4\}$ for illustration.
Given the critical value $\kappa(\mu_{0})$,
the solid line in gray (green) represents the RG running for $\lambda^{-1}$ without (with) $U(1)_X$ effect.
The RG runnings of $\kappa^{-1}$ (dotted) and $y^{-1}_{t}$ (dotdashed) shown in the figure imply that the upper bound $\kappa_{0}\leq 0.8$ is improved to $\kappa_{0}\leq 0.85\sim 0.86$ when the hidden $U(1)_{X}$ effect is taken into account.
Alternatively, given the same $\kappa(\mu_{0})$, the upper bound on $\lambda(\mu_{0})$ can be improved by the $U(1)_{X}$ effect.
Combination of the left and right panel shows that the deviation due to the change of $n_{5\bar{5}}$ is small, in comparison with the $U(1)_X$ effect.
Note that plots in Fig.1 are actually critical lines,
because the change from the perturbative to the non-perturbative region is rather abrupt.
\begin{widetext}
\begin{center}
\begin{figure}[ht]
\centering
\begin{minipage}[b]{0.5\textwidth}
\centering
\includegraphics[width=3in]{lowscale1.eps}
\end{minipage}%
\centering
\begin{minipage}[b]{0.5\textwidth}
\centering
\includegraphics[width=3in]{lowscale2.eps}
\end{minipage}%
\caption{RG running of $\lambda^{-1}$ (solid line) for initial EW value $\lambda(\mu_{0})=1.0$ and $g_{X}(\mu_{0})=0.82$ for $a=1/2$, and messenger parameters $M=10^{7}$ GeV and $n_{5\bar{5}}=1$ (left) and $n_{5\bar{5}}=4$ (right), respectively.
The gray and green color corresponds to $\lambda$-SUSY without and with $U(1)_{X}$, respectively.
The RG runnings of $\kappa^{-1}$ (dotted) and $y^{-1}_{t}$ (dotdashed) shown in the figure indicate that the upper bound $\kappa_{0}\leq 0.8$ is improved to $\kappa_{0}\leq 0.85\sim 0.86$ when the hidden $U(1)_{X}$ effect is taken into account.
Horizontal line refers to critical point between perturbative and non-perturbative region. }
\end{figure}
\end{center}
\end{widetext}
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$\lambda$(1TeV) & \ without $U(1)_X$ & \ with $U(1)_X$ \\
\hline\hline
1.0 & $\kappa(\mu_{0})\leq 0.80$ & $\kappa(\mu_{0})\leq 0.85$ \\
1.10 & $\kappa(\mu_{0})= 0$ & $\kappa(\mu_{0})\leq 0.81$\\
1.23& not well defined & $\kappa(\mu_{0})= 0$ \\
\hline
\end{tabular}
\caption{Upper bound on $\lambda(\mu_{0})$ for $g_{X}(\mu_{0})=0.82$ for $a=1/2$,
messenger parameters $M=10^7$ GeV and $n_{5\bar{5}}=1$, and different initial values of $\kappa(\mu_{0})$.
In contrast to the result $\lambda(\mu_{0})\leq 0.8$ for $\kappa(\mu_{0})\simeq 0$ in \cite{9801437},
we have $\lambda(\mu_{0})\leq 1.23$ instead for the case with the hidden $U(1)_{X}$.}
\end{center}
\end{table}
In table I, we show the improvement on the upper bound on $\lambda(\mu_{0})$
due to the hidden $U(1)_{X}$.
Without $U(1)_X$ sector, the model is not well defined up to the GUT scale when $\lambda >0.80$.
Instead, $\lambda\leq 1.23$ stays perturbative up to the GUT scale when the hidden $U(1)_X$ is added to the model.
The critical value on $\lambda$ may be modified due to different choices on parameters $\{n_{5\bar{5}}, M, a\}$.
As shown in Fig.2, the deviation to $\lambda(\mu_{0})$ due to messenger parameters $M$ and $n_{5\bar{5}}$ is rather small.
Similar conclusion holds when one tunes $a$.
Because $a$ larger than $1/2$ leads to larger contribution to $\delta\beta_{y_{t}}$ but
unfortunately smaller $g_{X}(\mu_{0})$, and vice versa.
In summary, in terms of introducing a hidden $U(1)_{X}$ sector with gauge
symmetry broken scale $\simeq 10$ TeV
well defined $\lambda$-SUSY up to the GUT scale is allowed for $\lambda\leq 1.23$.
In this class of models all fields including singlet $S$ and the Higgs doublets are fundamental.
The implication for $\lambda\sim 1-2$ has been addressed, e.g., in \cite{1310.0459}.
In the next section, we revise the phenomenological implication for such large $\lambda$ together with small $\kappa$.
\section{III.~PQ Symmetry and Small $\kappa$-Phenomenology}
This section is devoted to study the phenomenology in $\lambda$-SUSY with small $\kappa$.
Although it is a consequence of taking large $\lambda$ for well defined $\lambda$-SUSY,
the study of small $\kappa$-phenomenology can be considered as an independent subject from the viewpoint of phenomenology.
The smallness of $\kappa$ can be understood due to a broken $U(1)$ global symmetry,
i.e., Peccei-Quinn (PQ) symmetry.
Without $\kappa$, the model is invariant under the following $U(1)$ symmetry transformation,
\begin{eqnarray}{\label{PQ}}
H_{u}\rightarrow H_{u} \exp({i\phi}), ~H_{d}\rightarrow H_{d} \exp({i\phi}), ~
S\rightarrow S \exp({-2i\phi}), \nonumber\\
\end{eqnarray}
Terms like $\delta V=m^{2}S^{2}+B_{\mu} H_{u}H_{d}$ explicitly breaks this symmetry.
For these breaking small, we have a pseudo-Goldstone boson with small mass.
The soft terms in the scalar potential for small $\kappa$-phenomenology is given by,
\begin{eqnarray}{\label{soft}}
V_{soft}&=&m^{2}_{S}\mid S \mid ^{2}+m^{2}_{H_{u}} H^{2}_{u}+m^{2}_{H_{d}} H^{2}_{d}\nonumber\\
&+&(A_{\lambda}\lambda SH_{u}H_{d}+ \text{H.c}),
\end{eqnarray}
where $\kappa$ relevant terms are ignored.
The EW symmetry breaking vacuum can be determined from Eq.(\ref{soft}).
For details on analysis of EW breaking vacuum,
see, e.g., \cite{0910.1785, 0712.2903}.
There are five free parameters
\begin{eqnarray}
\{m^{2}_{H_{u}},~m^{2}_{H_{d}},m^{2}_{S}, ~A_{\lambda},~\lambda\}
\end{eqnarray}
which define the small $\kappa$-phenomenology.
The first two can be traded for $\upsilon=174$ GeV and $\tan\beta$.
In what follows we explore the constraints on these parameters,
the Higgs scalar spectrum, and their couplings to SM particles.
The $3\times3$ squared mass matrix of CP-even neutral scalars reads
\footnote{Here we follow the conventions and notation in Ref.\cite{0712.2903}. },
\begin{widetext}
\begin{eqnarray}{\label{matrix}}
\mathcal{M}^{2}=\left(%
\begin{array}{ccccc}
\frac{A_{\lambda}^{2}}{1+x}+(M^{2}_{Z}-\lambda^{2}\upsilon^{2})\sin^{2}2\beta & -\frac{1}{2}(M^{2}_{Z}-\lambda^{2}\upsilon^{2}) \sin4\beta& - A_{\lambda}\lambda\upsilon\cos 2\beta \\
* & M^{2}_{Z}\cos^{2}2\beta+\lambda^{2}\upsilon^{2}\sin^{2}2\beta & - A_{\lambda}\lambda\upsilon\sin 2\beta \frac{x}{1+x} \\
* & *& \lambda^{2}\upsilon^{2}(1+x) \\
\end{array}%
\right)
\end{eqnarray}
\end{widetext}
in the basis $(H, h, s)$,
where $x\equiv m^{2}_{S}/(\lambda\upsilon)^{2}$,
$H=\cos\beta h_{2}-\sin\beta h_{1}$ and $h=\cos\beta h_{1}+\sin\beta h_{2}$
under the compositions $H_{u}^{0}=\frac{1}{\sqrt{2}}(h_{1}+i\pi_{1})$,
$H_{d}^{0}=\frac{1}{\sqrt{2}}(h_{2}+i\pi_{2})$ and $S=\frac{1}{\sqrt{2}}(s+i\pi_{s})$.
The state $h$ mixes with other two states $H$ and $s$ via $\mathcal{M}^{2}_{12}$ and $\mathcal{M}^{2}_{23}$ , respectively.
Due to the smallness of $\mathcal{M}^{2}_{12}$,
state $h$ mixes dominately with $s$.
Without such mixing ( i.e., $\mathcal{M}^{2}_{23}\rightarrow 0$ ),
$h$ couples to SM gauge bosons and fermions exactly as the SM Higgs.
This implies that in the parameter space of small $m_{S}$,
$h$ is the SM-like scalar discovered at the LHC with $\mathcal{M}^{2}_{22}=(126 \ GeV)^{2}$,
and $s$ decouples from SM gauge bosons with mass $\lambda \upsilon\sim 209$ GeV for $\lambda=1.2$ .
Such scalar $s$ easily escapes searches performed by the LEP2 and Run-I of LHC.
The precision measurement of $h$ including its couplings to SM gauge bosons and fermions
powerfully constrains the magnitude of mixing effect.
After mixing, we define the mass eigenstates as $h_{1}$ and $h_{2}$,
where $h_{1}=\cos \theta h-\sin\theta s$, $h_{2}=\cos \theta s +\sin\theta h$.
Normalized to SM Higgs boson couplings,
$h_1$ and $h_2$ couple to SM particles as,
\begin{eqnarray}{\label{coupling}}
\xi_{h_{1}VV}&=&\frac{g^{2}_{h_{1}VV}}{g^{2}_{hVV}}=\cos^{2}\theta,~~
\xi_{h_{2}VV}=\frac{g^{2}_{h_{2}VV}}{g^{2}_{hVV}}=\sin^{2}\theta.\nonumber\\
\end{eqnarray}
where $V=\{W, Z, t, b,\cdots \}$.
At present status, LHC data suggests that $0.96\leq\cos^{2}\theta\leq 1$ \cite{fit}.
Alternatively we have $\sin^{2}\theta \leq 0.04$.
In Fig.3 we show the parameter space in the two-parameter plane of $m_S$ and $A_{\lambda}$,
for $\lambda=1.2$ and $\tan\beta=\{4.5, 5, 5.5\}$, respectively.
For each $\tan\beta$ region below the color contour is excluded by the condition of stability of potential
and mass bound on chargino mass $m_{\chi}>103$ GeV \cite{chargino}.
Region in the right up corner is excluded by the precision measurement of Higgs coupling presently.
In particular, $m_{S}$ heavier than $\sim$90, 105, and 110 GeV is excluded for $\tan\beta=4.5$, 5 and 5.5, respectively.
In comparison with the choice $\lambda=0.7$ discussed in \cite{0712.2903},
the main difference is that for $\lambda=1.2$ it allows larger $m_{S}$.
As for the Higgs mass constraint,
the discrepancy between $\mathcal{M}^{2}_{22}$ and 126 GeV is compensated by the stop induced loop correction.
The stop mass for the zero mixing effect (i.e., $A_t\sim 0$) is $\sim 340$ GeV for $\tan\beta=4.5$, $\sim 550$ GeV for $\tan\beta=5$ and $\sim 800$ GeV for $\tan\beta=5.5$.
Stop mass beneath 1 TeV is favored by naturalness.
Nevertheless, there has tension for such light stop with present LHC data.
This problem can be resolved in some situation,
which we will not discuss here.
We show in Fig.4 the ratio $\xi_{h_{2}VV}$ defined in Eq.(\ref{coupling}),
which determines the production rate for scalar $h_2$.
Its magnitude increases slowly as $m_S$ becomes larger.
$\xi_{h_{2}VV}$ reaches $\sim 0.03$ at most when $m_S$ closes to its upper bound (suggested by Fig. 3).
For such strength of coupling and mass $\sim 200$ GeV,
$h_2$ is easily out of reach of Run-I at the LHC and earlier attempts at the LEP2.
Small ratio of strength of coupling similarly holds for heaviest CP-even neutral state $H$.
In this sense, it is probably more efficient to probe charged Higgs scalar $H^{\pm}$, CP-odd scalar $A$, or light pseudo-boson $G$.
Studies along this line can be found in, e.g, \cite{0005308}.
\begin{figure}[ht]
\includegraphics[width=0.45\textwidth]{kappa.eps}
\caption{Parameter space in the two-parameter plane of $m_S$ and $A_{\lambda}$,
for $\lambda=1.2$ and $\tan\beta=4.5$ (black), 5 (blue), 5.5 (red), respectively.
For each $\tan\beta$ region below the color contour is excluded by the condition of stability of potential
and mass bound on chargino mass $m_{\chi}>103$ GeV.
Region in the right up corner is excluded by the fit to Higgs couplings presently.}
\end{figure}
\begin{figure}
\includegraphics[width=0.45\textwidth]{ratio.eps}
\caption{Ratio of coupling strength for $h_2$ shown in the plane of $m_{S}-A_{\lambda}$ for $\lambda=1.2$ and $\tan\beta=4.5$.}
\end{figure}
As for other scalar masses we show them in table II for three sets of $A_{\lambda}$ and $m_{S}$,
which are chosen in the parameter space shown in Fig.3.
It is shown that for each case the mass of charged Higgs boson exceeds its experimental bound $\sim 300$ GeV,
and $A$ scalar is always the heaviest with mass around 600 GeV.
In comparison with scalar mass spectrum in fat Higgs model \cite{0311349},
the spectrum in table II is similar to it,
although their high energy completions are rather different.
As a final note we want to mention that the smallness of $m_S$ in compared with $A_{\lambda}$ can be achieved in model building, e.g., in gauge mediation.
Because singlet $S$ only couples to messengers indirectly through Higgs doublets.
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c|c|}
\hline
$(A_{\lambda}, m_{S})$ (GeV) & $H^{\pm}$ (GeV) & $H$ (GeV)& $A$(GeV)\\
\hline\hline
(500,~40) & 463& 482& 535 \\
(560,~50) & 527 & 530 & 585 \\
(620,~50) & 590 & 586 & 640 \\
\hline
\end{tabular}
\caption{Mass spectrum for three sets of choices, which are subtracted from Fig.3.}
\end{center}
\end{table}
\section{IV.~Conclusions}
In this paper, we have studied $\lambda$-SUSY which stays perturbative up to the GUT scale.
We find that the bound $\lambda\leq0.7\sim 0.8$ in the minimal model
is relaxed to $\lambda\leq1.23$
if a hidden $U(1)_X$ gauge theory is introduced
above scale $\sim 10$ TeV
and small $\kappa(\mu_{0})$ is assumed at the same time.
This improvement gives rise to several interesting consequences in phenomenology.
For example, the fine tuning related to 126 Higgs mass can be reduced,
and light stop beneath 1 TeV can be allowed.
In the light of such single $U(1)_X$,
one may introduce multiple $U(1)$s or other gauge sectors at the intermediate scale,
and further uplift the bound on $\lambda$.
In the second part of the paper,
we have revised small $\kappa$-phenomenology for large $\lambda$.
In comparison with fat Higgs model \cite{0311349},
the spectrum in the small $\kappa$-phenomenology is similar,
although their high energy completions are different.
The null result for signals of the other two CP-even neutral scalars $h_2$ and $H$
is due to the perfect match between the scalar
discovered at the LHC (Here its is referred as $h_1$) and the SM Higgs.
Because the perfect fit dramatically reduces the mixing effect between $h_1$ and the others,
which results in tiny strength of coupling for $h_2$ and $H$ relative to the SM expectation.
The studies on signals of charged scalar $H^{\pm}$, CP-odd scalar $A$ and pseudo-boson $G$
will shed light on this type of model.\\
~~~~~~~~~~~~~~~~~~~
$\mathbf{Acknowledgement}$\\
The author thanks J-J. Cao for correspondence and the referee for valuable suggestions.
This work is supported in part by Natural Science Foundation of China under Grant No.11247031 and 11405015.
|
\section{Introduction}
\label{sec:introduction}
The presence of bars in galaxies is a feature that was noticed from
the earliest studies of external galaxies
\citep{Hubble-36,deVaucouleurs-63}. More recent estimates on the
fraction of barred disk galaxies fluctuate between one- to
two-thirds, depending on the method to identify bars and the inclusion
or not of weak bars
\citep{Mulchaey-97,Knapen-Shlosman-Peletier-00,Eskridge-00,
Marinova-Jogee-07,Sheth-08,Cameron-10, Lee-12a, Oh-Oh-Yi-12}. The
large abundance of stellar bars in local galaxies has triggered
interest on the involvement of these features on secular evolution
\citep{Sellwood-Wilkinson-93, Kormendy-Kennicutt-04, Cheung-E-13}. Given that
stellar bars break the radial symmetry of galaxies, they are
particularly efficient in redistributing matter and angular momentum
between different components, namely stars, gas and dark matter
\citep{Hohl-71,Sellwood-80,Tremaine-Weinberg-84,Weinberg-85,
Debattista-Sellwood-00,Athanassoula-02,Athanassoula-03,Martinez-Valpuesta-Shlosman-Heller-06}.
Clear examples of redistribution of material driven by bars is the
fuel of gas inward
\citep{Shlosman-Frank-Begelman-89,Friedli-Benz-93}, resulting in
accumulation of material that might be used as a build-up of
disk-like bulges or pseudo-bulges
\citep{Kormendy-Kennicutt-04,Athanassoula-05,Heller-Shlosman-Athanassoula-07}. Given
the collisional nature of gas, this component can effectively lose
energy during shocks and flow inward
\citep{Shlosman-Frank-Begelman-89,Friedli-Benz-93}, which help to
explain the high concentrations of molecular gas found on barred
galaxies \citep{Sakamoto-99,Sheth-05}, as well as the younger
stellar populations and higher star formation rates in the bulges of
barred galaxies when compared with unbarred ones
\citep[e.g.,][]{Coelho-Gadotti-11, Wang-12}. As a result, the overall
structure and morphology of barred galaxies gets shaped by the
effects of stellar bars.
At the same time, bar formation and evolution depends on many physical
properties of hosting galaxies. In general, bars are more frequently
found in massive, red galaxies with prominent bulges and overall
early-type morphologies \citep{Sheth-08,Weinzirl-09,
Hoyle-11,Masters-11,Lee-12a, Cheung-E-13}. Late-type, less massive, blue galaxies
also show the presence of stellar bars, but their frequency is low and
the bars they exhibit are shorter in size
\citep{Elmegreen-Elmegreen-85,Erwin-05,Aguerri-Gonzalez-Garca-09,Lee-12a}. \citet{Masters-12}
found that the bar fraction is significantly lower in gas-rich disk
galaxies than in gas-poor ones, a result in agreement with theoretical
expectations from \citet{Athanassoula-Machado-Rodionov-13}, where they
find in their simulations that bars form later in gas-rich systems,
and after they form, they grow more slowly than in poor-gas systems,
hence predicting that the bar likelihood should be higher in galaxies
with low gas content. Using numerical simulations of spinning dark
matter halos, \citet{Long-Shlosman-Heller-14} demonstrate that bar
growth in strength and size, is strongly suppressed on systems with
spin parameter $\lambda \gtrsim 0.03$, giving support to the
observational finding by \citet{Cervantes-Sodi-13}, where they reached
this same conclusion estimating $\lambda$ for the same galaxy sample
employed on the present study, and finding that strong bars are
confined to reside in galaxies with low to intermediate values of
$\lambda$, while weak bars are preferentially found in rapidly
spinning systems. More recently, \citet{Cervantes-Sodi-14} explored
the dependence of the bar fraction on the stellar-to-halo mass
ratio (M$_{\mathrm{*}}/$M$_{\mathrm{h}}$) for central disk galaxies finding,
in agreement with theoretical studies \citep{Ostriker-Peebles-73,
Efstathiou-82, Christodoulou-95, DeBuhr-12}, that bars are more common
in galaxies with high M$_{\mathrm{*}}/$M$_{\mathrm{h}}$ values, and
exploring the bar fraction in the M$_{\mathrm{h}}$ vs M$_{\mathrm{*}}$
plane, they find that the dependence is stronger considering a relation
with the form $f_{\mathrm{bar}}=f_{\mathrm{bar}}$(M$_{\mathrm{*}}^{\alpha}/$M$_{\mathrm{h}}$)
with $\alpha=1.5$.
Although secular evolution has proven to be an effective driver of
evolution, it is by no means the only way in which a galaxy can
transform. Regarding the formation and growth of bars, early
simulations of interacting disk galaxies by
\citet{Noguchi-87,Noguchi-88} show that self-gravitating disks that
were expected to be stable against the development of a bar, once
perturbed by the tidal force of a companion, develop prominent spiral
structures, that at the center resemble bar structures that are able
to induce gas infall into the nucleus when gas dissipates energy via
cloud-cloud collisions. Similar results are reported by
\citet{Byrd-Valtonen-90} for the case of tidal forces exerted by a
cluster potential on simulated disk galaxies. Although the
configuration of the interaction changes the final properties of the
bar, some groups found that in general, it enhances the formation of
the bar \citep{Gerin-Combes-Athanassoula-90,Miwa-Noguchi-98}, while
others report that the opposite is also possible: the bar can loose
angular momentum and mass and debilitate after the interaction
\citep{Sundin-Sundelius-91,Sundin-Donner-Sundelius-93}. There seem to
be also controversy about the different effects of prograde and
retrograde interactions
\citep{Romano-Daz-08,Aguerri-Gonzalez-Garca-09}.
Ultimately, the effect of environment on the formation and evolution
of bars needs to be addressed through observational evidence.
\citet{Thompson-81} found a larger fraction of barred galaxies in the
core of the Coma cluster and \citet{Giuricin-93} reported a higher
recurrence of early-type barred spirals in high local density
environments. Later studies of local \citep{Andersen-96,Eskridge-00}
and intermediate-redshift clusters \citep{Barazza-09} showed similar
findings and, more recently, \citet{Mendez-Abreu-12} proved that
by taking into account the different luminosity ranges, it is possible to
detect different effects of the environment for luminous and faint galaxies
located in clusters or in the field. They propose that interactions in
bright disk galaxies that are stable enough against interactions trigger
bar formation, while for faint galaxies, interactions are strong enough to
heat the disk inhibiting bar formation.
The opposite result--no evidence for a dependence of bar frequency
on environment, comparing field and cluster galaxies--is also widely
reported \citep{vandenBergh-02,
Marinova-09,Aguerri-Gonzalez-Garca-09,Mendez-Abreu-Sanchez-Janssen-Aguerri-10,
Giordano-11,Marinova-12}. When the samples include only S0 galaxies,
most of the time, a secondary dependence on environment is reported,
with a higher frequency of barred S0 galaxies in clusters
\citep{Barway-Wadadekar-Kembhavi-11} and an increase in the bar
fraction toward the cluster core \citep{Lansbury-Lucey-Smith-14}.
By computing projected redshift-space two-point cross-correlation
functions (2PCCF) for a sample of nearly 1,000 galaxies from the SDSS,
\citet[][hereafter L09]{Li-09} studied the clustering properties of
barred galaxies finding that the clustering of barred and unbarred
galaxies of similar stellar mass is indistinguishable and no evidence
that bar formation is promoted by mergers or interactions. More
recently, \citet{Skibba-12}, also using clustering statistics, reported
a positive correlation for bulge-dominated and barred galaxies to be
found in denser environments on scales of 150 kpc to a few megaparsecs than
their unbarred and disk-dominated counterparts and argued that barred
galaxies are often central galaxies in low-mass dark matter halos or
satellites in more massive ones.
\citet{Lee-12a} also investigated the dependence of the barred galaxy
fraction on environment, finding that once mass and color are fixed,
the bar fraction is independent of the background density, but an
influence of the nearest neighbor appears when the separation to the
nearest neighbor is less than 0.1 times the virial radius of the
neighbor. As the distance decreases, the bar fraction drops,
regardless of the morphology of the neighbor--a possible indication
that strong tidal interactions destroy bars or prevent their
growth. Similar results were presented by \citet{Mendez-Hernandez-11}
comparing reduced samples of isolated galaxies and galaxies in pairs,
showing that the subsample of galaxies in pairs presented a bar
fraction of only 20\%, while that for isolated systems was 43\%. Using
a much larger galaxy sample, \citet{Casteels-13} reported a similar
result using the Galaxy Zoo 2 sample, with a decrease for the
likelihood of identifying a bar in pairs with separations $\leq$30
$h^{-1}$ kpc.
In this paper, we study the environment of galaxies with bars in the
local universe, extending the analysis of L09 by using the new sample
provided by \citet{Lee-12a}, which is 30 times larger than the sample
used by L09. When compared to previous studies of the same topic, our
approach differs in the following ways.
\begin{itemize}
\item Instead of computing the auto-correlation of the barred galaxy
sample, we compute the cross-correlation of our sample with
respect to reference samples of the general galaxy population from
scales of a few tens of kiloparsecs up to a few tens of megaparsecs. This takes
advantage of the much larger sample size and volume of the
reference samples, allowing us to substantially increase the
signal-to-noise ratio in our results, as well as allowing us to study
the scale dependence of the clustering of barred galaxies in detail.
\item We wish to isolate the link between environment and the
presence or absence of a bar in galaxies, so the environment
measurements of the barred galaxies are always compared to those
of control samples of unbarred galaxies that are closely matched
in galaxy properties known to be correlated with
environment. According to previous studies of the clustering of
galaxies \citep[e.g.][]{Li-06a}, we consider stellar mass,
optical color and internal structure, quantified by the surface
stellar mass density and the concentration of the stellar light
distribution, when constructing the control samples.
\item In addition to computing the 2PCCFs, we use three other
statistics to characterize the environment of our galaxies. These
include (1) the background-subtracted average neighbor counts
($N_c$) around the barred and unbarred galaxies as a function of
the projected distance to neighboring galaxies, (2) the overdensity
($\delta$) of the local environment of our galaxies estimated on
$\sim3$ Mpc scale, and (3) the membership of our galaxies in the SDSS
group systems identified by \citet{Yang-07}. The different
statistics have their pros and cons in quantifying galaxy
environment, and provide measurements that are complementary to
each other.
\end{itemize}
This paper is organized as follows. Section 2 gives a brief
description of the volume-limited sample, as well as the control and
reference samples used in this study. In Section 3, we describe the
methods used to characterize the environment of the galaxies in our
sample. In Section 4, we present our general results. Finally, in
Section 5 we summarize our results and present our conclusions.
Throughout this paper, we assume a cosmology with a density parameter of
$\Omega = 0.3$, cosmological constant of $\Omega_\Lambda =0.7$, and
Hubble constant of $H_0$ = 70 km s$^{-1}$ Mpc$^{-1}$.
\section{Data}
\subsection{The sample of barred galaxies}
\label{sec:data}
The sample of galaxies with bar identification used in our study comes
from a previous work by \citet{Lee-12a}. Here we give a brief
description of the sample selection and morphology classification, and
the reader is referred to \citet{Lee-12a} for the full description of
the sample and comparisons with previous classifications
\citep{deVaucouleurs-91,Nair-Abraham-10} and Park \& Lee (2014) for the
publicly available catalog. The sample is a
volume-limited sample selected from the Korea Institute for Advanced
Study Value-Added Galaxy Catalog, which was constructed from the
SDSS DR7 \citep{Abazajian-09} by \citet{Choi-Han-Kim-10} and consists
of 33,391 galaxies with $r$-band absolute magnitudes brighter than
$M_{r} =$ --19.5 + 5log$h$ and spectroscopically measured redshift in
the range 0.02 $\leq z \leq$ 0.05489. The segregation into early- and
late-type galaxies is done by adopting the prescription by
\citet{Park-Choi-05}, where galaxies are divided according to their
morphology in color versus color gradient and concentration index
space plus an additional visual inspection. The identification of
bars is done by visual inspection of $g+r+i$ combined color images
from SDSS. Early-type galaxies are classified as either barred or
unbarred galaxies. For the case of late-type galaxies, given that the
classification is more robust for face-on systems, we limit our sample
to galaxies with a minor-to-major axis ratio of $b/a>0.6$. For
late-type galaxies presenting a bar, the bar is further classified as
{\em strong} or {\em weak} according to its size--if the bar is larger
than one quarter the size of the hosting galaxy, it is classified as a
strong bar, otherwise it becomes a weak bar.
Given that the bar fraction estimated by visual inspection of
optical images is statistically the same as that estimated using
automated methods such as ellipse fitting \citep{Menendez-Delmestre-07,Sheth-08}
and that the computation of the bar fraction gives the same result using
near-infrared images as it does when using optical \citep{Whyte-02,Menendez-Delmestre-07,
Sheth-08}, we are confident that our results are reproducible even if
the bar detection is obtained by using a different method.
We extract the physical quantities required for our work from two
additional catalogs: the New York University Value Added Galactic
Catalog \citep[NYU-VAGC;][]{Blanton-05}
\footnote{http://sdss.physics.nyu.edu/vagc/} and the MPA/JHU SDSS
database \citep{Kauffmann-03, Brinchmann-04}
\footnote{http://www.mpa-garching.mpg.de/SDSS/}, both of which are
based on SDSS DR7 and publicly available. These include stellar mass
$M_\ast$, stellar surface mass density $\mu_\ast$, color $(g-r)$ and
concentration index $C$. The stellar mass of a galaxy was estimated
based on its redshift and the five-band SDSS photometry following the
methodology detailed in \citet{Kauffmann-03}
\footnote{Note that the masses included in the current MPA/JHU
database are based on fits to photometry and so are not identical to
those of \citet{Kauffmann-03} where the masses are obtained from
spectroscopic indices. See
http://www.mpa-garching.mpg.de/SDSS/DR7/mass\_comp.html for detailed
comparisons between the two masses.}. The surface stellar mass
density of a galaxy is defined by $\mu_\ast=0.5M_\ast/\pi R_{50,z}^2$,
where $R_{50,z}$ is the radius enclosing 50\% of the $z$-band flux, and
the concentration index $C=R_{90,r}/R_{50,r}$, where $R_{90,r}$ and
$R_{50,r}$ are the radii enclosing 90\% and 50\% of the total light in
the $r$-band image of the galaxy. Our final sample includes 17,839
unbarred and 3,749 barred galaxies. In order to distinguish this
sample from subsequent control samples, we will refer to it as C0
sample, which we will divide into eight different subsamples: AU (AB)
includes all unbarred (barred) galaxies, EU (EB) includes only
unbarred (barred) early-type galaxies, LU (LB) includes only
unbarred (barred) late-type galaxies, SB includes only strongly barred
late-type galaxies and WB includes weakly barred late-type galaxies.
\subsection{Control samples of unbarred galaxies}
Previous studies have shown that barred and unbarred galaxies present
different stellar mass, color, and surface mass density
distributions,and at the same time the clustering of galaxies strongly
depends on all these properties \citep[e.g.][]{Li-06a}.In order to
isolate the effect of environment on bars, we construct a series of
control samples of unbarred galaxies, and in what follows we compare
our results of each barred-galaxy subsample always with those of the
corresponding control sample. We compile two different control
samples: C2 encompasses galaxies with matched stellar mass and color
between barred and unbarred galaxies and C3 matches an extra
quantity--the stellar surface mass density. The matching tolerances
are $\triangle$log$M_* \leq$ 0.08 , $\triangle(g-r)\leq$ 0.025,
$\triangle$log$\mu_*\leq$ 0.08.
\begin{figure*}
\begin{center}
\epsfig{figure=1-eps-converted-to.pdf,width=\hsize}
\caption{Projected 2PCCF for barred $w_p^{bar} (r_p)$ and
unbarred $w_p^{unbar} (r_p)$ galaxies (top panels) and
$w_p^{bar} (r_p)$ to $w_p^{unbar} (r_p)$ ratio (bottom
panels) for the different subsamples of our parent galaxy sample
C0. Panels from left to right correspond to the whole galaxy
sample of barred (AB orange triangles) plus unbarred (AU black
diamonds) galaxies, early-types barred (EB orange triangles) and
unbarred (EU black diamonds), late-types barred (LB orange
triangles) and unbarred (LU black diamonds), strongly barred
late-types (SB yellow triangles) and LU, and weakly barred
late-types (WB blue triangles) and LU.}
\label{fig:2PCCF_C0}
\end{center}
\end{figure*}
\subsection{Reference galaxy samples}
We have constructed two reference samples from NYU-VAGC: a {\em
spectroscopic} reference sample, which is used to compute the
projected two-point cross-correlation function(2PCCF) $w_p(r_p)$, and
a {\em photometric} reference sample, which is used to calculate
counts of close neighbors around our barred and unbarred galaxies.
The spectroscopic reference sample is constructed from version {\tt
dr72} of the NYU-VAGC, the current version built on the SDSS DR7
data, by selecting all galaxies with $r$-band apparent magnitude
corrected for foreground extinction $r<17.6$ within the restricted
redshift range of $0.01\leq z \leq 0.2$ and the absolute magnitude
range $-23\leq M_{0.1_r} \leq -17$. Here $M_{0.1_r}$ is the $r$-band
absolute magnitude corrected to its value at $z = 0.1$. With these
selection criteria, we are left with about half a million galaxies in
our spectroscopic reference sample. The photometric reference sample
is also constructed from NYU-VAGC version {\tt dr72} by selecting
galaxies with $r$-band apparent magnitude down to $r<21$. The sample
includes about 2.5 million galaxies.
\section{Methodology}
In this section we briefly describe the methods we will use to
characterize the environment of the galaxies in our samples. For a
full description of these methods, we refer the reader to the source
papers where they are presented in detail
\citep{Li-06a,Li-06b,Yang-07,Jasche-10}.
\subsection{Projected two-point cross-correlation function}
We use the projected redshift-space
2PCCF, $w_p(r_p)$, to quantify the clustering properties of
our galaxies. For this we have constructed a random sample that has
the same selection effects as the spectroscopic reference sample
following the method described in \citet{Li-06a}. We cross-correlate
each of our samples of barred galaxies, as well as the corresponding
control sample of unbarred galaxies, with respect to both the
spectroscopic reference sample and the random sample, and then define
$w_p(r_p)$ as a function of the projected separation $r_p$ by the
ratio of the two pair counts minus one. A careful correction for the
effect of fiber collisions is implemented by comparing the angular
2PCF of the spectroscopic sample with that of the parent photometric
sample \citep[see][for details]{Li-06b}. Errors of the $w_p(r_p)$
measurements are estimated using the bootstrap resampling technique
\citep{Barrow-Bhavsar-Sonoda-84}, for which a total of 100 bootstrap
samples are generated based on the real sample. Details about our
methodology for computing the correlation functions and for
constructing the random sample can be found in \citet{Li-06a}.
\subsection{Close neighbor counts}
We count the number of galaxies in the photometric reference sample in
the vicinity of the galaxies in our samples, and we make a statistical
correction for the effect of chance projections by subtracting the
average count around randomly placed galaxies. When compared to the
2PCCF, close neighbor counts are not affected by incompleteness on
small scales caused by the fiber collisions. In addition, since the
photometric sample is much deeper than the spectroscopic sample, by
computing background-subtracted neighbor counts in the photometric
sample one is able to include much fainter companion galaxies, thus
probing the effect of close companions over a broader range of mass
ratios.
\begin{figure*}[t]
\begin{center}
\epsfig{figure=2-eps-converted-to.pdf,width=\hsize}
\end{center}
\caption{Same as the Figure 1, but using control sample C2 where
the galaxies share a common distribution of stellar mass $M_*$
and color g$-$r. }
\label{fig:2PCCF_C2}
\end{figure*}
\subsection{Group catalog}
To account for the environment of galaxies in groups, we make use of
the group catalog of \citet{Yang-07}, constructed from the sample {\tt
dr72} of the NYU-VAGC. The groups of galaxies are identified using
a modified version of the halo-based group-finding algorithm developed
in \citet{Yang-05} and applied to a sample with redshifts in the range
0.01 $\leq z \leq$ 0.20 and with a redshift completeness greater than
0.7. From this extensive catalog, we use only galaxies identified in
groups with at least three members, from which we are left with a
final sample of nearly 2,000 galaxies. For each group system, we take
the most massive galaxy as the central galaxy, and the rest as
satellites.
\subsection{Large-scale overdensity}
To account for the environment at large scales , we make use of a
non-linear, non-Gaussian, full Bayesian large-scale structure analysis
of the cosmic density field based on the SDSS DR7 by
\citet{Jasche-10}, where the authors use a Bayesian sampling algorithm
in order to obtain the extremely high dimensional non-Gaussian,
non-linear log-normal Poissonian posterior of the three-dimensional
density field. The reconstruction of the field is made over a 750 Mpc
cube, taking into account the angular and radial selection functions
of the SDSS, and a proper treatment of a Poissonian shot noise
contribution for an effective resolution of the order of $\sim$3 Mpc.
\section{Results}
\subsection{Cross-correlation functions}
\begin{figure*}[t]
\begin{center}
\epsfig{figure=3-eps-converted-to.pdf,width=\hsize}
\end{center}
\caption{ Same as the Figure 1, but using control sample C3 where
the galaxies share a common distribution of stellar mass $M_*$,
color g$-$r, and stellar surface mass density $\mu_*$. }
\label{fig:2PCCF_C3}
\end{figure*}
\begin{figure*}[t]
\begin{center}
\epsfig{figure=4-eps-converted-to.pdf,width=\hsize}
\end{center}
\caption{Histograms for stellar mass $M_*$, $g-r$ color, stellar
surface mass density $\mu_*$ and the concentration index C for
control samples C0 (dotted lines) and C3 (solid lines). The first
line corresponds to the whole sample, the second line to early-types,
and the third line to late-types.}
\label{fig:sample_properties}
\end{figure*}
\begin{figure}[t]
\begin{center}
\epsfig{figure=5-eps-converted-to.pdf,width=\hsize}
\end{center}
\caption{Same as the Figure 1, but using control sample C3 for
early-type galaxies with 2.8$\leq C \leq$ 3.5. }
\label{fig:2PCCF_con}
\end{figure}
\begin{figure*}[]
\begin{center}
\epsfig{figure=6-eps-converted-to.pdf,width=\textwidth}
\end{center}
\caption{Average counts of galaxies in the photometric sample
within a given projected radius R$_P$ from the galaxies in the
C3 subsamples. Each line corresponds to different apparent
magnitude limits in the $r$ band ($r\leq$ 18,19,20,20.5,21) for the
galaxies in the photometric sample.}
\label{fig:nc}
\end{figure*}
In Figure~\ref{fig:2PCCF_C0}, we show the 2PCCF $w_p(r_p)$ measured
for our main galaxy sample C0. In each panel we present $w_p(r)$ for
barred $w_p(r)^{bar}$ and unbarred $w_p(r)^{unbar}$ galaxies for the
different subsamples in consideration (from left to right: all the
galaxies in the sample, early-types, late-types, strongly barred
late-types and weakly barred late-types). For clarity we also include
the amplitude ratio $w_p(r)^{bar}$ to $w_p(r)^{unbar}$ for each case.
At first sight, it is clear that the clustering of our barred and
unbarred samples are very similar for all the subsamples, although
some differences appear when looking carefully at each case. Starting
with the extreme left panel, which includes all the galaxies in our
sample, we notice that unbarred galaxies are more strongly clustered
than barred ones on scales below $\sim$1 Mpc. If we include only
early-types in our analysis (second panel), the clustering ratio
between barred and unbarred systems changes, presenting a pick at
around 1 Mpc and positive values at larger scales. Only on scales
bellow 100 kpc does the ratio becomes less than unity. When we turn to
the case of late-type galaxies (middle panel), barred galaxies appear
to be more strongly clustered than their unbarred counterparts at
almost all the scales probed. If we distinguish between strong (fourth
panel) and weak (fifth panel) bars, we see that the signal seen for
the case of late-type galaxies mostly comes from the clustering of
strongly barred galaxies, which is natural given that they are more
numerous in our sample than weakly barred galaxies.
At this point, Figure~\ref{fig:2PCCF_C0} seems to indicate that the
clustering of barred and unbarred galaxies is different, but we need
to be careful to drive conclusions at this stage given the following
two facts. First, previous studies of galaxy clustering have well
established that clustering strongly depends on galaxy properties
including mass, color and morphological type \citep[e.g.][]{Li-06a,
Zehavi-11}. Second, it is also well known that the presence or
absence of bars is not the only morphological difference between our
barred and unbarred samples. For instance, previous works
\citep{Barazza-Jogee-Marinova-08,Nair-Abraham-10,Lee-12a} show that
the fraction of barred galaxies is higher in massive red galaxies
than in less-massive blue systems. Massive red galaxies are more
strongly clustered than less-massive, blue galaxies. To normalize out
this effect we make use of our control samples C2 and C3. As
described in Section 2.1, for the case of C2 we have
restricted our samples so that each subsample of barred galaxies has
the same distribution in stellar mass and color as the corresponding
subsample of unbarred galaxies, while for the case C3 the subsamples
of barred and unbarred galaxies being compared are additionally required
to have the same distribution in stellar surface mass
density.
Measurements of the 2PCCF for our C2 and C3 samples are shown in
Figures ~\ref{fig:2PCCF_C2} and~\ref{fig:2PCCF_C3}, respectively. For
late-type galaxies, the different clustering behaviors between barred
and unbarred galaxies seen in Figure~\ref{fig:2PCCF_C0} disappear when
the samples being compared are closely matched in stellar mass, color
and surface mass density, with the ratio $w_p(r_p)^{bar}$ to
$w_p(r_p)^{unbar}$ consistent with unity within error bars on all
scales probed. The same result holds if we further split our sample of
barred late-type galaxies into strongly barred and weakly barred
galaxies according to the size of their bars. This is consistent with
the expectation that the relatively strong clustering of late-type,
barred galaxies seen for the full sample (i.e. sample C0) is indeed
attributed to the fact that the subset of barred galaxies with
late-type morphology in C0 are more massive and redder, thus more
strongly clustered, when compared to the unbarred late-type galaxies.
This result implies that the presence of a bar in late-type galaxies
is not obviously linked with their clustering properties as quantified
by the 2PCCF, confirming the earlier finding of \citet{Li-09} that was
based on a much smaller sample.
\begin{figure*}[t]
\begin{center}
\epsfig{figure=7-eps-converted-to.pdf,width=\textwidth}
\end{center}
\caption{Ratio of average neighbor counts of barred to unbarred
galaxies for C3 within a given apparent magnitude limit in the
$r$ band ($r\leq$ 18,19,20,20.5,21). }
\label{fig:nc_ratio}
\end{figure*}
For early-type galaxies, in contrast, the difference in the 2PCCF
between barred and unbarred galaxies is still significantly seen for
scales larger than a few hundred kiloparsecs, even when the samples are closely
matched in mass, color, and surface mass density. Considering that the
clustering of galaxies depends not exclusively on these parameters, we
compare the distributions of
stellar mass, g$-$r color, surface mass density, and concentration
index $C$ for samples C0 and C3 in Figure~\ref{fig:sample_properties}. In the case of C3, the distributions
for $M_*$, g$-$r and $\mu_*$ are indistinguishable between barred and
unbarred galaxies, regardless of their morphological type, but the
concentration index $C$ follows different statistical distributions
between barred and unbarred systems, especially for the case or
early-type galaxies (see the rightmost panels). In order to further
isolate the link between bars and the clustering of early-type
galaxies, we remove the dependence of clustering on concentration
index from our analysis by restricting ourselves to a narrow range of
concentration index 2.8$\leq C \leq$ 3.5, and we repeat the same
analysis for early-type galaxies as we did above for the case of C3.
The result is presented in Figure~\ref{fig:2PCCF_con}. Although the
difference on large scales ($>1$Mpc) becomes less significant, the
stronger clustering of barred galaxies still persists on intermediate
scales, at around 500 kpc. As shown in \citet{Li-08a,Li-08b}, on these
scales the 2PCCF is dominated by pairs of galaxies hosted by the same
dark matter halo. Thus the difference in 2PCCF between barred
early-type galaxies and unbarred early-type galaxies implies that the
two types of galaxies are distributed in their host dark matter halo
in different ways. We will come back to this point in Section 4.4.
\subsection{Close neighbor counts}
A problem with computing the 2PCCF on very small scales using our
spectroscopic sample is the effect of fiber collisions. Although we
implemented a statistical correction that is proved correct on
average, it might still introduce systematic effects in our study. In
this subsection we investigate the clustering of our samples by
computing the background-subtracted neighbor counts, $N_c(r_p<R_p)$,
which is by definition the number of galaxies in the photometric
reference sample within the projected radius $R_p$ of the barred or
unbarred galaxies, with the effect of chance projection being
statistically corrected (see Section 2). When compared to the 2PCCF
analyzed above, this quantity doesn't suffer from the fiber collision
effect on small scales. It also allows us to explore the effect of
fainter companion galaxies. For this analysis we only consider our C3
samples where the dependence of clustering on mass, color and internal
structure of galaxies has been taken into account.
In Figures~\ref{fig:nc} and~\ref{fig:nc_ratio}, we plot $N_C$,
measured within a given value of
projected radius $R_P$, for barred and unbarred galaxies in sample
C3. As in the previous subsection, we make comparisons for barred and
unbarred galaxies for given morphological type with their corresponding ratios of
$N_C$ for barred galaxies relative to those of unbarred
ones. Figure~\ref{fig:nc}, top to bottom, shows the results for the whole
sample, the early-type, and late-type subsamples, respectively, and in Figure~\ref{fig:nc_ratio},
we split present again the result for late-types as well as the result from
splitting the sample into strong and weak bars. Panels in both figures, from left to
right, show the results for different apparent $r$-band limiting
magnitude applied to the photometric reference sample ranging from
$r_{lim}=18$ for the leftmost panels to $r_{lim}=21$ for the
rightmost panels.
In Figure~\ref{fig:nc} (top panels), where we compare the neighbor counts
around all the barred galaxies and the control sample of unbarred
galaxies, we see that barred galaxies have less companions on scales of
$\la$50 kpc than unbarred galaxies, while on scales of $\ga$100 kpc,
barred galaxies present more neighbors than their unbarred
counterparts. When we split the galaxies according to their
morphological type, we find that the differences on the small and
large scales, found in the top panels for the barred galaxies as a whole,
are essentially contributed by late-type and early-type galaxies
separately. As the second row of panels show, the higher neighbor
count on large scales is an effect seen only for early-type galaxies,
indicating that barred, early-type galaxies are located in higher
density regions than unbarred, early-type galaxies of similar mass,
color and internal structure. This confirms what we find above from
the 2PCCF measurements, where the same sample of barred galaxies shows
higher clustering amplitude on intermediate scales than the control
sample of unbarred galaxies. When the comparison is limited to
late-type galaxies, the difference on large scales is no longer
significant, while at the smallest scales we see clearly the drop in
$N_c$, suggesting that the same effect found for the full sample is
driven by late-type galaxies.
Figure~\ref{fig:nc_ratio} presents the result for late-type galaxies only,
making the distinction between strong and weak bars.
The large number of strongly barred galaxies in our sample helps to
improve the statistics and reduce the statistical errors in our
measurements, but even with large error bars, if we look at the result
for weakly barred galaxies we see that the drop of average neighbor
counts is noticeable at even larger scales, up to a few hundred kiloparsecs. A
possible explanation for this effect at larger scales for the case of
weakly barred galaxies might be related with these systems being less
massive than those galaxies hosting strong bars \citep{Lee-12b,
Cervantes-Sodi-13}, which would make them less stable against
interactions with massive neighbors even at large separations.
Finally, we notice that in all the cases we have considered, the
differences in the neighbor count change very little as one goes to
fainter and fainter limiting magnitudes. This implies that the excess
or drop in neighbor counts seen above is contributed primarily by
relatively bright companions with $r<18$, but not the fainter
companions.
\subsection{Galaxy group catalog}
Our results studying the 2PCCF and average number counts for our
sample of galaxies indicate that barred early-type galaxies
present a higher clustering amplitude and a larger number of companion
galaxies on scales from a few hundred kiloparsecs to several megaparsecs when compared
with unbarred ones. In particular, Figure~\ref{fig:2PCCF_con} reveals
a maximum at $\sim500$kpc in the ratio of the 2PCCF between barred and
unbarred early-type galaxies, with the ratio decreasing at both
smaller and larger scales. This implies that the 2PCCFs of the two
samples are different not only in amplitude, but also in shape on these
scales. As shown in \citet{Li-06b}, the amplitude of 2PCCF on scales
larger than a few megaparsecs is a direct measure of the mass of host dark
matter halos, while the shape of the 2PCCF on a few $\times100$kpc is
sensitive to the relative fraction of central and satellite galaxies
in the galaxy sample. Our result here thus suggests that early-type
barred galaxies are expected to be found more frequently in satellite
galaxies than in central galaxies.
To further explore this possibility we make use of the group catalog
of \citet{Yang-07} in order to identify central and satellite galaxies
in our sample. As described in the previous section, after matching
our main sample with the group catalog and only keeping galaxies in
groups with three or more members, our final sample reduces to nearly
2,000 galaxies. From this reduced sample we find that the ratio of
central-to-satellite systems for barred and unbarred late-type
galaxies is very similar, about 27\% (see also \citet{Cervantes-Sodi-14}).
For early-type galaxies this
ratio is significantly different: 24\% for unbarred and only 16\% for
barred galaxies. Although the small sample size doesn't allow us to
do a more detailed analysis, the different central-to-satellite galaxy
number ratio for early-type galaxies is already telling us that
early-type barred galaxies in our sample are indeed more commonly
found as satellites in galaxy groups, reinforcing our hypothesis from
the results using 2PCCFs and neighbor counts.
\subsection{Large-scale overdensity}
\begin{figure*}
\begin{center}
\epsfig{figure=8-eps-converted-to.pdf,width=\textwidth}
\end{center}
\caption{Cumulative distribution of barred (color line) and
unbarred (black line) galaxies as a function of the overdensity
parameter $\delta$ for the different subsamples of C3.}
\label{fig:overdensity}
\end{figure*}
Finally, we have also examined the overdensity of the large-scale
environment of our barred and unbarred galaxies. For this, we use the
overdensity $\delta$ estimated over a scale of $\sim3$Mpc by
\citet{Jasche-10} through a reconstruction of the three-dimensional
density field in the local universe based on the SDSS DR7
spectroscopic galaxy sample, as described on Section 3.4. The results
are shown in Figure~\ref{fig:overdensity}, where we compare the
cumulative distributions of the different subsamples of our control
sample C3 as a function of ln(2 + $\delta$). As can be seen, the
distributions of barred and unbarred galaxies are almost identical in
all cases. We conclude that the presence of a bar in galaxies is
unlikely correlated with the overdensity of local environment on
scales of $\sim3$Mpc. This is in good agreement with what we find
above from the 2PCCFs and close neighbor counts on large scales.
\section{Summary and discussion}
We have presented an extensive study of the environment of galaxies
with bars in the low-redshift universe, which makes use of a
volume-limited sample of $\sim$ 30,000 galaxies from the SDSS with
visually determined morphological classifications and bar
identifications. We use four distinct statistics to characterize the
environment of the barred galaxies in our sample: the
2PCCF $w_p(r_p)$ of our galaxies with
respect to a spectroscopic reference sample, the background-subtracted
average counts $N_c(R_p)$ of neighboring galaxies in a photometric
reference sample, the membership of our galaxies in the SDSS group
systems, and the overdensity $\delta$ of the local environment at
$\sim3$ Mpc scale. We segregate our galaxies into early- and
late-types to study if differences arise between the two different
morphological types. For the late-type subsample, we further
distinguish two types of barred galaxies, those with strong and weak bars,
according to their size. We apply the four statistics to the
subsamples of barred galaxies, and compare the results to the
same statistics obtained for control samples of unbarred galaxies that
are closely matched to the barred galaxy samples in stellar mass,
color, and internal structure in order to normalize out the dependence
of environment on these parameters.
The first thing we notice measuring $w_p(r_p)$, once we remove any
dependence on stellar mass, color and stellar surface mass density,
is that the clustering for barred and unbarred galaxies are very
similar, with only minor differences. The differences arise when we
split the galaxies in our sample according to their morphological
types. For the case of late-type galaxies, the 2PCCF does not depend
on the presence or absence of a bar, confirming the previous result
by \citet{Li-09}. The barred early-type subsample shows a higher
amplitude for $w_p(r)$ on scales from a few hundred kiloparsecs to 1 Mpc, when
compared with unbarred galaxies, very similar to the result reported
by \citet{Skibba-12}, who found a higher likelihood for barred and
bulge-dominated galaxies on denser environments on scales of 150 kpc
to a few megaparsecs, than unbarred, disk-dominated ones. At these
intermediate scales, the correlation function is dominated by the
one-halo term, which would indicate that barred early-type galaxies
are more frequently found as satellite systems. To explore this idea
further, we made use of the \citet{Yang-07} group catalog to identify
our galaxies in groups as centrals or satellites, finding that the
percentage of barred early-type galaxies was significantly higher for
satellites compare with the percentage for centrals. In contrast, the
percentage of barred late-type galaxies is the same for central as
for satellite galaxies, in agreement with recent results by \citet{Cervantes-Sodi-14}.
With our barred early-type sample composed by S0 galaxies, this
result is also in agreement with \citet{Barway-Wadadekar-Kembhavi-11}
who found that S0 galaxies in clusters present a higher bar fraction
than their counterparts in the field, and
\citet{Lansbury-Lucey-Smith-14}, who analyzed a reduced sample of S0
galaxies in Coma and found an increase in the bar fraction from the
outskirts of the cluster toward the core. A possible framework to
explain this trend is that these S0 galaxies were originally spirals
that were transformed by ram pressure stripping with the intracluster
medium, loosing their cold interstellar gas to the environment,
causing a fading of the disk and subsequently a morphology change, but
preserving their stellar features such as the bar. This mechanism is
expected to work only for faint lenticular galaxies
\citep{Boselli-Gavazzi-06,Barway-09}, which are not able to retain
their gas, while bright S0s presumably form through other
processes (e.g., mergers). Interestingly, our barred lenticulars are
mostly faint galaxies, with M$_r > -21.5$. With a higher bar
fraction for spirals than for lenticulars \citep{Laurikainen-09},
this could explain the higher amplitude found for the correlation
function in our sample of barred early-type galaxies on scales that
correspond to satellite systems of more massive halos.
Given that our estimate for the 2PCCF suffers the effect of fiber
collisions at small scales, we turned to close neighbor counts to
study the clustering of barred galaxies at small scales. In this
case, it is the barred late-type subsample that shows a
systematic trend of having an average lower number of neighbors on
scales of $\le$50 kpc when compared with the control sample of unbarred
late-type galaxies. The effect seems more dramatic for the case of
weakly barred galaxies, where the deficiency appears at larger
scales close to 100 kpc. This finding may indicate that tidal
interactions destroy and/or disfavor bar growth instead of
enhancing it as expected from results of numerical simulations. As
weak bars are more commonly found in fainter, less-massive galaxies
than strong bars \citep{Lee-12b,Cervantes-Sodi-13}, the fact that we
find the effect of nearby companions at larger scales for the case of
weak bars may be not unexpected: due to their lower mass their
gravitational potential is also weaker and they are more fragile to
disturbances than their more massive counterparts. In addition, the
decline of the average close neighbor count around the barred
galaxies in our sample is found to be weakly dependent on the
limiting magnitude of the photometric sample, implying that the
effect is primarily driven by relatively bright companion galaxies,
with little contribution from companions fainter than $r=18$.
Our result echoes previous studies that report a decrease on the bar
fraction for disk galaxies when the separation to the nearest
neighbor becomes smaller than 0.1 times the virial radius of the
neighbor \citep{Lee-12a}, a lower fraction of barred galaxies in
close pairs when compared with isolated galaxies
\citep{Mendez-Hernandez-11} and a decrease on the likelihood of
identifying a bar in pairs with small separations
\citep{Casteels-13}. In any case, we only find indications of
suppressed formation/growth of stellar bars in galaxies that might be
undergoing some kind of interaction with a close neighbor, with no
evidence of environmental stimulation, in opposition with earlier
studies \citep{Elmegreen-Elmegreen-Bellin-90}.
Finally, we do not find any dependence of the likelihood of galaxies
hosting bars on the large-scale environment, as accounted by the
overdensity parameter $\delta$, estimated through a reconstruction of
the three-dimensional cosmic web posterior \citep{Jasche-10}. This result suggests
that galactic bars are not obviously linked to the large-scale
structure of the universe.
\acknowledgments
The authors thank the anonymous referee for comments that improved the paper.
We are grateful to Changbom Park for providing the galaxy sample used
in this study and Ramin Skibba, Edmond Cheung, and Jairo Mendez-Abreu
for helpful discussions.
This work is sponsored by NSFC (grant Nos. 11173045, 11233005, 11325314,
11320101002) and the Strategic Priority Research Program ``The Emergence of
Cosmological Structures'' of CAS (grant No. XDB09000000).
Funding for the SDSS and SDSS-II has been provided by the Alfred P.
Sloan Foundation, the Participating Institutions, the National Science
Foundation, the U.S. Department of Energy, the National Aeronautics
and Space Administration, the Japanese Monbukagakusho, the Max Planck
Society, and the Higher Education Funding Council for England. The
SDSS Web site is http://www.sdss.org/. The SDSS is managed by the
Astrophysical Research Consortium for the Participating
Institutions. The Participating Institutions are the American Museum
of Natural History, Astrophysical Institute Potsdam, University of
Basel, University of Cambridge, Case Western Reserve University,
University of Chicago, Drexel University, Fermilab, the Institute for
Advanced Study, the Japan Participation Group, Johns Hopkins
University, the Joint Institute for Nuclear Astrophysics, the Kavli
Institute for Particle Astrophysics and Cosmology, the Korean
Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos
National Laboratory, the Max-Planck-Institute for Astronomy (MPIA),
the Max-Planck-Institute for Astrophysics (MPA), New Mexico State
University, Ohio State University, University of Pittsburgh,
University of Portsmouth, Princeton University, the United States
Naval Observatory, and the University of Washington.
|
\section{I. INTRODUCTION}
Bose--Einstein condensates (BECs) of dilute atomic gases are an invaluable
source of insight into the quantum behavior of matter at macroscopic scales
\cite{PeSmB,PiStB}. Topics like vortex structures or nonlinear matter waves
are being elucidated through BEC investigations \cite{Carre1}. In particular,
considerable effort has been invested in recent years in the study of solitary
matter waves (solitons) \cite{RevDS,Kon1,Mor1} due, in part, to their
similarity with well-known equivalent structures in the field of nonlinear
optics \cite{KivGOP,Kiv2D,RMP-Malomed}. A great variety of solitons have been
experimentally observed on the background of condensed atomic clouds. Dark
solitons have been found \cite{Burger1,Phil1,Hau1,Corn1,DS-2} as excitations
in systems with repulsive interatomic interactions (the analogue of
self-defocusing nonlinear media), while bright solitons have been generated
\cite{Hul1,Salo1,Wie1} as ground states when the interaction is attractive
(self-focusing nonlinear media). A particularly important special case of
bright solitons is that of gap solitons \cite{Mor1}. These matter-wave
nonlinear structures are localized states that can occur in BECs with
repulsive interatomic interactions if a periodic potential (such as that
produced by an optical lattice) is present. Although gap solitons have been
observed and studied in nonlinear optics since some time ago, only more
recently have they been found in BEC experiments \cite{Ober1}. This has
triggered a renewed theoretical interest in the study of these nonlinear
structures.
Strictly speaking, solitons in a one-dimensional (1D) setting are exact
solutions of the one-dimensional nonlinear Schr\"{o}dinger equation
(1D NLSE).
Accordingly, in the framework of the mean-field approximation, matter-wave
solitonic structures in 1D optical lattices have usually been investigated
theoretically by means of the 1D Gross--Pitaevskii equation (GPE)
\cite{GS1D,Kiv1,Efre1,Abd1,Mal1,Wu1}, which is nothing but a nonlinear
Schr\"{o}dinger equation describing the evolution of the condensate
wavefunction in the presence of the confining potentials. However,
soliton-like states realized in BECs always have an intrinsic
three-dimensional (3D) character, even when the trapping potential imposes
a tight radial confinement. In fact, typical experimental parameters for the
generation of gap solitons in BECs lie in a region of parameter space where
the quasi-1D approximation is not strictly valid \cite{Mor1,Ober1} and, as a
consequence, very often the 1D GPE cannot quantitatively reproduce the
dynamics of realistic matter-wave gap solitons.
Various approaches have been put forward with the aim of incorporating in the
description of condensed gases in elongated geometries the consequences of
their intrinsic 3D character. Effective models for studying the stationary
properties of highly anisotropic BECs beyond the analytically solvable
perturbative and Thomas--Fermi regimes were proposed in Refs. \cite{Das1,VDB1,VDB2}.
An analytic expression for the local chemical potential $\mu_{\bot}\!(n_{1}\!)$ of
elongated condensates as a function of the axial linear density was independently
proposed in Refs. \cite{VDB2} and \cite{Fuch1}.
Such expression remains valid in the 3D--1D crossover regime and, in the case of
\cite{VDB2}, is also valid in the presence of an axially symmetric vortex.
Effective equations have also been introduced to study topics such as solitons
\cite{Jack1}, the influence of the transverse confining on the dynamics of BECs
in 1D optical lattices \cite{Yuka1}, or the emergence of Faraday waves
\cite{Nico3}. Of particular relevance because of their simplicity and usefulness are
the effective 1D equations derived in Refs. \cite{Salas1} and \cite{Ef1DEqs,VDB2},
which govern the condensate dynamics accounting properly for the contributions from
excited transversal modes through a nonpolynomial nonlinear term. Several variants
of the above equations aimed at improving their accuracy to the price of
incorporating adjustable \cite{Adhik1} or a larger number of variational parameters
\cite{Kam1,Clark1} have also been proposed. However, the intended increase in
accuracy hardly compensates for the complexity of the resulting equations or the
arbitrariness in the choice of free parameters.
On the other hand, a number of generalizations of the above equations have been
derived to study issues such as the expansion of elongated BECs \cite{Modug1},
spin-1 atomic condensates \cite{You1}, matter-wave vortices \cite{Salas3}, BECs in
anisotropic transverse traps \cite{Salas4}, binary condensates in mixed dimensions
\cite{Adhik2}, dark--bright solitons in two-component BECs \cite{Carre3}, dipolar
BECs \cite{Adhik4}, or localized modes in condensates with spin-orbit and Rabi
couplings \cite{Salas6}.
The equations of Refs. \cite{Salas1} and \cite{Ef1DEqs} have also been used to
study BECs in nonlinear lattices \cite{Salas5}, bright solitons in condensates with
inhomogeneous defocusing nonlinearities \cite{Mal2}, or the Bogoliubov spectrum of
elongated condensates \cite{Ben1}.
Since these effective equations are extensions of the standard
1D GPE (to which they reduce in the proper limit) and can be numerically
solved with similar computational effort, they may offer a clear advantage
over the latter. In this work, we will focus on the effective equation
proposed in Refs. \cite{Ef1DEqs,VDB2}. This equation has been used in the
description of matter-wave dark solitons in elongated BECs and has proven to
give results in good agreement with the experiments
\cite{Kev1,Wel1,Kev2,Isa1}. As was demonstrated in Ref. \cite{VDB7}, it can
also be applied to the study of stationary gap solitons in the different
physically relevant mean-field regimes.
This is a nontrivial issue because this kind of systems are characterized by a
small spatial scale (the lattice period $d$) that can represent a rather
restrictive condition. However, to establish the utility of the model in
realistic situations and determine whether it may be a real alternative to a
fully 3D treatment one has to analyze its ability to reproduce a number of
physical properties of fundamental interest in the theoretical description of
stationary matter-wave gap solitons that were not studied in Ref. \cite{VDB7}.
One of these properties is the dependence of the soliton chemical potential
$\mu$ on the number $N$ of constituent atoms, a quantity that can be used to
define trajectories in the $\mu-N$ plane which are essential for the
classification of gap solitons into different families. Another one is the
frequency spectrum of the elementary excitations determining the dynamical
stability of the gap soliton family. To provide an answer to the previous
question, in this work we analyze to what extent the 1D GPE and the effective
equation of Refs. \cite{Ef1DEqs,VDB2} can reproduce the above-mentioned
fundamental physical properties in realistic situations. To this end we
consider condensed atomic clouds confined by transverse harmonic potentials
in a range of parameters typical of current experiments and compare with
exact numerical results obtained from the full 3D GPE.
Since in this work we are interested in the mean-field regime, we shall
restrict our analysis to physical configurations with a large number of atoms
per lattice well, which are amenable to a treatment in terms of a mean-field
macroscopic wavefunction satisfying the GPE.
The paper is organized as follows. Section II briefly reviews the theoretical
models considered in this work. In Secs. III and IV we analyze the numerical
results: in Sec. III we focus on the wavefunctions and the distinctive
trajectories of the different gap soliton families, and in Sec. IV we study
the frequency spectrum of the elementary excitations and the stability
properties of these nonlinear structures. Finally, in Sec. V we summarize our
results and present our conclusions.
\section{II. EFFECTIVE MODELS FOR GAP SOLITONS IN 1D OPTICAL LATTICES}
When a periodic potential is imposed on a BEC, the chemical potential of the
system develops a band-gap structure. In these circumstances, it is possible
to find either extended states with chemical potential inside the bands
(nonlinear Bloch waves) or localized gap solitons between bands. At zero
temperature and in the mean-field regime, these coherent states are accurately
described by the 3D GPE \cite{PeSmB,PiStB}:
\begin{equation}
i\hbar\frac{\partial\psi}{\partial t}=\left( -\frac{\hbar^{2}}{2m}\nabla
^{2}+V(\mathbf{r})+gN\left\vert \psi\right\vert ^{2}\right) \psi,
\label{3DGPE}%
\end{equation}
where $g=4\pi\hbar^{2}a/m$ is the interaction parameter, with $a$ being the
\textsl{s}-wave scattering length, $N$ is the number of atoms, and
$V(\mathbf{r})$ is the potential of the trap. In this work we will consider an
external potential of the form $V(\mathbf{r})=V_{\bot}(r_{\bot})+V_{z}(z)$,
which besides the axial term $V_{z}(z)=V_{0}\sin^{2}(\pi z/d)$ accounting for
a 1D optical lattice with period $d$, includes a transverse harmonic trapping
$V_{\bot}(r_{\bot})=m\omega_{\bot}^{2}r_{\bot}^{2}/2$, where $r_{\bot}%
^{2}=x^{2}+y^{2}$.
A description in terms of the GPE corresponds to a classical mean-field
description that neglects quantum correlations. It represents the classical
limit of the underlying quantum field theory and becomes an excellent
approximation in the limit $N \rightarrow \infty$ with the nonlinear
interaction parameter $gN$ held constant.
In practice, quantum fluctuations decay rapidly with $N$ and one expects the
GPE to be a good approximation for configurations with a few hundreds of atoms
per lattice well.
Although the 3D GPE can be solved numerically with no
approximations \cite{PRL06,Mott}, it demands a considerable computational
effort when either the 3D character of the system or the nonlinearity
increases. In order to simplify the computational problem, simpler models with
reduced dimensionality are commonly used in the study of gap solitons in 1D
optical lattices. Whenever it is possible to make the assumption that the
condensate remains in the ground state of the transverse harmonic trap (so
that higher transverse modes are not excited), a good approximation to Eq.
(\ref{3DGPE}) is given by the 1D GPE with rescaled nonlinearity:
\begin{equation}
i\hbar\frac{\partial\phi}{\partial t}=-\frac{\hbar^{2}}{2m}\frac{\partial
^{2}\phi}{\partial z^{2}}+V_{z}(z)\phi+g_{1D}N\left\vert \phi\right\vert
^{2}\phi, \label{1DGPE}%
\end{equation}
where $g_{1D}=2a\hbar\omega_{\bot}$ is the 1D interaction parameter, obtained
after averaging over the `frozen' transverse degrees of freedom, and
$\phi(z)$ is the axial wavefunction of
the condensate. It is important to bear in mind that this equation is only
valid when the axial energies are small enough in comparison with the radial
quantum, $\hbar\omega_{\bot}$, a condition that, in general, is not satisfied
in realistic situations.
A simple dimensional analysis shows that the physical problem is characterized
by three dimensionless parameters which can be conveniently chosen as%
\begin{equation}
\frac{E_{R}}{\hbar\omega_{\bot}},\;\frac{V_{0}}{E_{R}},\;\frac{Na}%
{d} \label{con1}%
\end{equation}
where $E_{R}\equiv\hbar^{2}k^{2}/2m$ is the recoil energy of the lattice, with
$k=\pi/d$ being the lattice wavevector, $s=V_{0}/E_{R}$ is the lattice depth,
and $Na/d$ is the nonlinear coupling constant determining the mean-field
interaction energy. The derivation of Eq. (\ref{1DGPE}) relies on the
assumption that the 3D wavefunction $\psi$ can be factorized in independent
radial and axial components, with the radial component corresponding to the
Gaussian wavefunction of the ground state. Such assumption entails an implicit
adiabatic approximation according to which the transverse degrees of freedom
can instantaneously follow the axial dynamics. In order for this to be true,
the characteristic time scale of the axial motion must be much longer than
that of the radial motion ($\sim\omega_{\bot}^{-1}$). Moreover, since the
validity of the 1D GPE (\ref{1DGPE}) requires the radial configuration to
remain in the ground state, the axial energies involved must be sufficiently
small in comparison with the characteristic energy of the transverse harmonic
oscillator, $\hbar\omega_{\bot}$. Taking into account that the axial energy of
the underlying linear problem is of the order of the recoil energy $E_{R}$ and
that the nonlinear interaction energy is of the order of $an_{1}\hbar
\omega_{\bot}$ (with $n_{1}=N/d$ being the linear density of the gap soliton
along the axial direction), the above requirements can be stated as%
\begin{equation}
E_{R}\ll\hbar\omega_{\bot},\;\;an_{1}\ll1.\label{con2}%
\end{equation}
These conditions, which ensure the permanence of the system in the transverse
ground state, also guarantee automatically the validity of the adiabatic
approximation. However, in order for the mean-field approximation implicit in
Eqs. (\ref{3DGPE}) and (\ref{1DGPE}) to be applicable, it is also necessary
that $N\gg1$. Incorporating this latter requirement, the conditions for the
validity of the 1D GPE can be rewritten as%
\begin{equation}
d\gg a_{\bot},\;\;1\ll N\ll d/a,\label{con3}%
\end{equation}
where $a_{\bot}=\sqrt{\hbar/m\omega_{\bot}}$ is the radial harmonic-oscillator
length. These conditions are very restrictive and usually are not met in real
condensates. For instance, for $^{87}$Rb atoms $a=5.3$ $%
\operatorname{nm}%
$ and $a_{\bot}=10.78/\sqrt{\omega_{\bot}/2\pi}$ $%
\operatorname{\mu m}%
$ (with $\omega_{\bot}/2\pi$ given in $%
\operatorname{Hz}%
$) while, typically, the lattice parameter $d$ varies between $0.4$ and $1.6$
$%
\operatorname{\mu m}%
$ \cite{Arim1}, the lattice depth $s$ between $0$ and $30$ \cite{Ingus1} and,
for elongated condensates, $\omega_{\bot}/2\pi$ varies between $85$ and $1000$
$%
\operatorname{Hz}%
$. Thus, typically, $d\lesssim1.6$ $%
\operatorname{\mu m}%
$, $a_{\bot}\gtrsim0.34$ $%
\operatorname{\mu m}%
$, and $d/a\lesssim300$ which makes the conditions (\ref{con3}) hard to
satisfy. In particular, in the experiment of Ref. \cite{Ober1}, where gap
solitons were observed in a $^{87}$Rb condensate, $d=0.39$ $%
\operatorname{\mu m}%
$, $a_{\bot}=1.17$ $%
\operatorname{\mu m}%
$, $s=0.7$, and $N\sim900$ atoms. In that of Ref. \cite{Ingus1}, $d=0.42$ $%
\operatorname{\mu m}%
$, $a_{\bot}=1.14$ $%
\operatorname{\mu m}%
$, $s=0-30$, and $N\sim3\times10^{5}$ $^{87}$Rb\ atoms. It is important to
note, however, that even though most experiments have been carried out in the
above parameter regimes, larger values for $d$ and $\omega_{\bot}$ (which can
realize the 1D mean-field regime) are well within experimental reach.
Next, we briefly recall the 1D effective model proposed in
Refs. \cite{Ef1DEqs,VDB2}, which is an extension of the standard 1D GPE.
As before, the starting point is the adiabatic approximation, which allows the
factorization of the 3D wavefunction in the form%
\begin{equation}
\psi(\mathbf{r},t)=\varphi(\mathbf{r}_{\bot};n_{1})\phi(z,t), \label{II-4}%
\end{equation}
where the radial wavefunction $\varphi$ is assumed to be normalized to unity
and $n_{1}$ is the local linear density%
\begin{equation}
n_{1}(z,t)=N|\phi(z,t)|^{2}=N\int d^{2}\mathbf{r}_{\bot}|\psi(\mathbf{r}%
_{\bot},z,t)|^{2}. \label{II-5}%
\end{equation}
After substituting Eq. (\ref{II-4}) in Eq. (\ref{3DGPE}) and averaging over
the radial degrees of freedom one obtains the following 1D effective equation%
\begin{equation}
i\hbar\frac{\partial\phi}{\partial t}\!=\!-\frac{\hbar^{2}}{2m}\frac
{\partial^{2}\phi}{\partial z^{2}}+V_{z}\phi+\mu_{\bot}\phi, \label{II-6}%
\end{equation}
where $\mu_{\bot}\!(n_{1}\!)$ is the transverse local chemical potential which
follows from the stationary 2D GPE%
\begin{equation}
\left( \!-\frac{\hbar^{2}}{2m}\nabla_{\bot}^{2}+V_{\bot}(\mathbf{r}_{\bot
})+gn_{1}\left\vert \varphi\right\vert ^{2}\!\right) \varphi=\mu_{\bot}%
(n_{1})\varphi. \label{II-7}%
\end{equation}
When the dimensionless linear density is small enough ($an_{1}\ll1$), this
latter equation can be solved perturbatively and, to the lowest order
(\textit{i.e.}, for $\varphi(\mathbf{r}_{\bot})$ given by the Gaussian ground
state of the corresponding harmonic oscillator), one obtains $\mu_{\bot}%
=\hbar\omega_{\bot}(1+2an_{1})$. Substituting back in Eq. (\ref{II-6}) and
taking into account Eq. (\ref{II-5}) it is thus clear that one arrives at the
1D GPE (\ref{1DGPE}). However, a more general solution to Eq. (\ref{II-7}) can
be found. Using two independent approaches it was demonstrated in Refs.
\cite{VDB2,Ef1DEqs} that, for condensates with no vorticity (the case we are
interested in here), a very accurate solution, valid for arbitrary $an_{1}$,
is given by%
\begin{equation}
\mu_{\bot}=\hbar\omega_{\bot}\sqrt{1+4an_{1}}. \label{II-8}%
\end{equation}
Substituting this expression in Eq. (\ref{II-6}), one finally finds%
\begin{equation}
i\hbar\frac{\partial\phi}{\partial t}=-\frac{\hbar^{2}}{2m}\frac{\partial
^{2}\phi}{\partial z^{2}}+V_{z}(z)\phi+\hbar\omega_{\bot}\sqrt{1+4aN\left\vert
\phi\right\vert ^{2}}\phi. \label{E1DE}%
\end{equation}
This effective 1D equation governs the axial dynamics of the condensate,
having a much wider range of applicability than the standard 1D GPE
(\ref{1DGPE}). In particular, now the linear density $an_{1}$ can take
arbitrary values and in the limit $an_{1}\ll1$ one recovers Eq. (\ref{1DGPE}).
Yet, the applicability of the above equation still requires the fulfillment of
the adiabatic approximation and, as far as the dynamical problem is concerned,
this condition can be hard to satisfy in the presence of an optical lattice.
Things are different, however, for the stationary problem, which is, in fact,
the most relevant in the characterization of matter-wave gap solitons. In this
case, the time scale of the axial dynamics tends to infinity and the adiabatic
approximation always holds. As a result, unlike the usual 1D GPE
(\ref{1DGPE}), the effective model (\ref{E1DE}) will be able to correctly
predict the trajectories in $\mu-N$ plane as well as the wavefunctions and
stability properties of the different gap soliton families.
\begin{figure}[t]
\begin{center}
\includegraphics[width=7.3cm]{Fig1.jpg}
\end{center}
\caption{(Color online) The family of fundamental gap solitons as predicted by
the different 1D effective models. $N$ represents the number of $^{87}$Rb
atoms, and $\tilde{\mu}\equiv(\mu-\hbar\omega_{\bot})/E_{R}$ is the
dimensionless chemical potential. The solid red curve is the prediction from
the 1D effective equation (\ref{E1DE}), the dashed blue curve is that from the
1D GPE (\ref{1DGPE}), and the open symbols are exact results obtained by
solving numerically the full 3D GPE (\ref{3DGPE}). The inset shows a
representative example of a gap soliton in this family (point A) in terms of a
density isosurface (corresponding to $5\%$ of its maximum value). The local
phase (which is uniform in this case) is also represented as a color map.}%
\label{Fig1}%
\end{figure}
\section{III. GAP SOLITON FAMILIES}
Next, we will apply both Eqs. (\ref{1DGPE}) and (\ref{E1DE}) to the study of
stationary gap solitons in a regime corresponding to typical experimental
parameters. To this end, by using a Newton continuation method we look for
numerical solutions of the form $\phi(z,t)=\phi_{0}(z)e^{-i\mu t/\hbar}$, with
$\mu$ being the condensate chemical potential. These solutions define
distinctive trajectories in $\mu-N$ plane which, as already said, are
essential for the classification of gap solitons into different families. To
investigate how accurately the above 1D equations can predict such
trajectories and, thus, the physical properties of the various families, we
compare the predictions from both equations with those obtained from the
numerical solution of the full 3D GPE.
In what follows we consider a $^{87}$Rb condensate subject to a rather tight
transverse potential of frequency $\omega_{\bot}/2\pi=800$ $%
\operatorname{Hz}%
$ and in the presence of a 1D optical lattice with period $d=1.5%
\operatorname{\mu m}%
$ and depth\ $s=$ $20$. For these parameters, the recoil energy of the lattice
takes the value $E_{R}=0.32\hbar\omega_{\bot}$.
\begin{figure}[t]
\begin{center}
\includegraphics[width=8.5cm]{Fig2.jpg}
\end{center}
\caption{(Color online) Axial densities (left panels) and wavefunctions (right
panels) of the fundamental gap solitons corresponding to points A and B in
Fig. \ref{Fig1}. Solid red lines are the predictions of the 1D effective
equation (\ref{E1DE}), dashed blue lines are those of the 1D GPE
(\ref{1DGPE}), and open symbols are exact results obtained by solving
numerically the full 3D GPE (\ref{3DGPE}).}%
\label{Fig2}%
\end{figure}
Figure \ref{Fig1}\ shows the different theoretical predictions the above
equations make for the family of fundamental gap solitons. This figure depicts
the dimensionless chemical potential $\tilde{\mu}\equiv(\mu-\hbar\omega_{\bot
})/E_{R}$ of the solitons versus the number $N$ of constituent atoms. The
3D band-gap structure of the underlying linear problem is shown
in the background, with shaded stripes corresponding to the allowed energy
bands. The two narrow stripes located at $\tilde{\mu}=10.45$ and $16.7$ are
purely 3D Bloch bands associated with excited transversal states and, thus,
cannot be accounted for by any 1D model. They are replicas of the lowest
energy band (located at $\tilde{\mu}=4.2$) shifted up in energy, respectively,
by two and four quanta of the radial harmonic oscillator, $\hbar\omega_{\bot
}/E_{R}$ \cite{GS-3D,VDB8}. Open symbols in this figure are exact results
obtained by solving numerically the full 3D GPE (\ref{3DGPE}), the solid red
curve is the prediction from the 1D effective equation (\ref{E1DE}), and the
dashed blue curve is that from the standard 1D GPE (\ref{1DGPE}). As is
evident, this latter equation clearly fails. In fact, even in the first (1D)
gap, it makes errors larger than $100\%$ and thus can hardly be considered
valid. On the contrary, the effective model of Eq. (\ref{E1DE}) is in good
quantitative agreement with the\ 3D results over the entire curve. Gap
solitons in this family are nonlinear stationary states exhibiting a single
peak well localized at a lattice site. A representative example is displayed
in the inset, which shows the 3D atom density of a gap soliton of this family
containing $325$ particles (point A marked by an open triangle in the figure).
Its chemical potential, $\mu=4.7$ $\hbar\omega_{\bot}>\hbar\omega_{\bot}$,
reflects the fact that the corresponding transversal state is composed of
multiple harmonic-oscillator modes. The prediction the effective model
(\ref{E1DE}) makes for this soliton has an error in the number of particles of
only $2\%$. Figure \ref{Fig1} also shows that while the results from the 1D
GPE get worse as the number of atoms increase, those from the effective
equation (\ref{E1DE}) remain in good quantitative agreement with the 3D results.
\begin{figure}[t]
\begin{center}
\includegraphics[width=7.5cm]{Fig3.jpg}
\end{center}
\caption{(Color online) Same as Fig. \ref{Fig1}, but for the symmetric
composite family of gap solitons consisting of two in-phase peaks.}%
\label{Fig3}%
\end{figure}
This behavior is corroborated by Fig. \ref{Fig2}. Panels A and B in this
figure display the dimensionless axial densities $an_{1}$ (left panels) and
wavefunctions $\tilde{\phi}(z)=\sqrt{aN}\phi_{0}(z)$ (right panels) of the
fundamental gap solitons corresponding, respectively, to points A and B in
Fig. \ref{Fig1}. As before, solid red curves have been obtained from the
effective equation (\ref{E1DE}), dashed blue curves from the standard 1D GPE,
and open symbols are numerical results obtained from the full 3D GPE. For
reference, the lattice potential is also shown in arbitrary units (dotted
lines). In order to deduce the axial wavefunction from the corresponding 3D
results we have used that $\tilde{\phi}(z)=\sqrt{an_{1}(z)}e^{i\varphi(z)}$,
where the linear density can be readily obtained by performing the integral in
Eq. (\ref{II-5}). As for the axial phase $\varphi(z)$, we have defined it as
the average local phase at every $z$-plane. Again the agreement between the
predictions of Eq. (\ref{E1DE}) and the 3D results is very good, while the 1D
GPE clearly fails.
Similar conclusions can be drawn in the case of multiple gap solitons. Figure
\ref{Fig3} displays the trajectories in ${\mu}-N$ plane corresponding to the
composite family consisting of two in-phase peaks (see also Fig. \ref{Fig4}).
Solitons in this family are bound states formed by a symmetric superposition
of two fundamental gap solitons of the family $(1,0,0)$, \textit{i.e.} those
displayed in Fig. \ref{Fig1}. Here we are using the notation introduced in
Ref. \cite{GS-3D}, which enables us to classify the different fundamental gap
solitons into families characterized by the quantum numbers $(n,m,n_{r})$
corresponding to the 3D Bloch band from which they bifurcate, with
$n=1,2,3,\ldots$ being the band index of the corresponding 1D axial problem
and $m=0,\pm1,\pm2,\ldots$ and $n_{r}=0,1,2,\ldots$ being, respectively, the
angular-momentum and radial quantum numbers characterizing the transversal
state. In this work, as is usually the case, we restrict ourselves to gap
solitons consisting of fundamental constituents of the type $(n,0,0)$, which
are amenable to an effective treatment in terms of the 1D equations
(\ref{1DGPE}) and (\ref{E1DE}). Gap solitons with $(m,n_{r})\neq(0,0)$ have
been rarely considered in the literature \cite{GS-3D,VDB8}. They feature a
nontrivial radial topology and require a more elaborate treatment \cite{VDB8}.
The inset in Fig. \ref{Fig3} shows the atomic density (as an isosurface taken
at $5\%$ of the maximum density) of the soliton corresponding to point D
(marked by an open triangle). This soliton contains $1765$ atoms and has a
chemical potential $\tilde{\mu}=16.6$. As can be seen, the two fundamental
constituents occupy adjacent lattice sites and have the same phase (shown in
the inset in terms of a color map). This is also apparent from Fig.
\ref{Fig4}, which depicts the axial linear densities $an_{1}$ and
wavefunctions $\tilde{\phi}(z)$ of the gap solitons marked by C and D in Fig.
\ref{Fig3}. A simple comparison also shows that the wavefunctions of the
composite solitons are indeed symmetric superpositions of\ those in Fig.
\ref{Fig2}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=8.5cm]{Fig4.jpg}
\end{center}
\caption{(Color online) Same as Fig. \ref{Fig2}, but for the gap solitons
corresponding to points C and D in Fig. \ref{Fig3}.}%
\label{Fig4}%
\end{figure}
Gap solitons consisting of two out-of-phase peaks are analyzed in Fig.
\ref{Fig5}. The upper trajectories in this figure (those bifurcating from the
first Bloch band) correspond to composite solitons formed by an antisymmetric
superposition of two\ fundamental $(1,0,0)$ gap solitons. An example of a
soliton of this family is displayed in the upper inset, which shows a
representation of the 3D wavefunction of the soliton corresponding to point E
in the figure. This wavefunction is given in terms of an isosurface of the
atomic density (taken at $5\%$ of its maximum) and a color map indicating the
local phase at each point. As in the previous case, the two fundamental
constituents of this compound soliton occupy adjacent lattice sites, but now
they exhibit a $\pi$ jump in their relative phase. A detailed comparison of
the axial density and wavefunction of this soliton (given in the upper panels
of Fig. \ref{Fig6}) with those of the corresponding fundamental solitons
(given in the lower panels of Fig. \ref{Fig2}) confirms that this soliton is a
bound state consisting of the antisymmetric superposition of two of
the\ fundamental gap solitons previously studied.
\begin{figure}[t]
\begin{center}
\includegraphics[width=7.5cm]{Fig5.jpg}
\end{center}
\caption{(Color online) Same as Fig. \ref{Fig1}, but for the antisymmetric
composite (upper curves) and fundamental (lower curves) gap soliton families
consisting of two out-of-phase peaks.}%
\label{Fig5}%
\end{figure}
The lower trajectories in Fig. \ref{Fig5}, which bifurcate from the second 1D
Bloch band, constitute a family that has been referred to as the
\emph{subfundamental} family \cite{Mal1}. Gap solitons in this family are
associated with the first excited ($n=2$) energy band of the corresponding 1D
axial problem, which reflects in the fact that they exhibit two major peaks in
the axial direction (\textit{i.e.}, one axial node) localized in a single
lattice site (see the lower inset in Fig. \ref{Fig5} as well as panels F in
Fig. \ref{Fig6}). Unlike the previous case, these solitons (which can be
regarded as excited states featuring an embedded dark soliton) are not bound
states of two fundamental $(1,0,0)$ gap solitons. They are fundamental
solitons of the $(2,0,0)$ family which can be used as elementary constituents
of more complex compound structures.
\begin{figure}[t]
\begin{center}
\includegraphics[width=8.5cm]{Fig6.jpg}
\end{center}
\caption{(Color online) Same as Fig. \ref{Fig2}, but for the gap solitons
corresponding to points E and F in Fig. \ref{Fig5}.}%
\label{Fig6}%
\end{figure}
As is apparent from the above results, the predictions from the 1D effective
model (\ref{E1DE}) are always in good quantitative agreement with the 3D
numerical results.
\section{IV. STABILITY ANALYSIS}
Stability is of primary importance for the experimental relevance of gap
solitons. In this Section we are interested in determining whether the
effective equation (\ref{E1DE}) is able to correctly predict the stability
properties of the gap solitons previously considered. To this end we perform
a stability analysis based on the above effective equation and compare with
the respective 3D results. We begin by deriving the Bogoliubov--de Gennes
(BdG) equations corresponding to Eq. (\ref{E1DE}) and then obtain
numerically the frequency spectrum of the elementary excitations.
To derive the BdG equations, we perturb a given gap soliton stationary
wavefunction $\phi_{0}(z)$ in the form \cite{Carre1}
\begin{equation}
\phi(z,t)=\left[ \phi_{0}(z)+u(z)e^{-i\omega t}+v^{*}(z)e^{i\omega t}\right]
e^{-i\mu t/\hbar} \label{BDG1}%
\end{equation}
and substitute in Eq. (\ref{E1DE}) retaining only linear terms in the
amplitudes $u$ and $v$ of the normal modes of the system. The numerical
solution of the resulting linear equations provides the frequencies of the
elementary excitations determining the dynamical stability of the soliton. A
similar procedure has been followed to derive the BdG equations corresponding
to the 3D GPE (\ref{3DGPE}).
Results for gap solitons B, D, E, and F are given in the panels labeled with
the same letters in Fig. \ref{Fig7}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=8.5cm]{Fig7.jpg}
\end{center}
\caption{(Color online) Panels B, D, E, and F show the frequency spectra of
the elementary excitations of the gap solitons labeled with the same letters
as follow from the BdG equations corresponding to the 1D effective model
(\ref{E1DE}) (left panels) and to the 3D GPE (right panels). The horizontal
axis in each figure represents the real part of the excitation frequencies
$\omega$ shifted by the chemical potential $\tilde{\mu}_{0}$ of the
corresponding stationary soliton, $\tilde{\mu}=\tilde{\mu}_{0}%
+\mathrm{\operatorname{Re}}(\omega/E_{R})$, while the vertical axes show the
respective imaginary parts, $\mathrm{\operatorname{Im}}(\omega/\omega_{\bot}%
)$. Note that in all cases considered $\tilde{\mu}_{0}=16.6$.}%
\label{Fig7}%
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=8.5cm]{Fig8.jpg}
\end{center}
\caption{(Color online) Representative examples showing the long time
evolution of the stationary solitons D, E, and F after a random small
perturbation (see text for details).}%
\label{Fig8}%
\end{figure}
Left panels show the corresponding excitation spectra as obtained from the 1D
effective model (\ref{E1DE}), while right panels display the eigenfrequencies
of the BdG equations associated to the 3D GPE. The horizontal axis in each
figure represents the real part of the excitation frequencies $\omega$ shifted
by the chemical potential $\tilde{\mu}_{0}$ of the corresponding stationary
soliton, $\tilde{\mu}=\tilde{\mu}_{0}+\mathrm{\operatorname{Re}}(\omega
/E_{R})$. Vertical axes show the respective imaginary parts,
$\mathrm{\operatorname{Im}}(\omega/\omega_{\bot})$. As is well known, the
absence of complex eigenfrequencies in the spectrum implies the linear
stability of the system. Of course, the excitation spectrum of the full 3D
problem necessarily contains frequencies that cannot exist in the effective
problem with reduced dimensionality (for instance, those associated to radial
normal modes). This is an expected result. The point, however, is that, except
in those cases where the solitons exhibit a nontrivial (unstable) transverse
configuration, one might also reasonably expect that the stability properties
of the different gap solitons will be dictated by the axial degrees of
freedom, which ultimately are the ones responsible for the existence of
self-trapping. If this expectation proves true, the effective 1D model
(\ref{E1DE}) could be a reliable tool for the investigation of the stability
of these nonlinear structures. As is apparent from a comparison between left
and right panels in Fig. \ref{Fig7}, this is indeed the case: except for the
gap soliton F (lower panels), which is a bright soliton containing a dark
soliton that extends considerably in the radial direction, the effective 1D
model predicts accurately the complex eigenfrequencies determining the
stability of the various gap solitons. According to the excitation spectra
shown in the figure, the fundamental soliton B and the symmetric composite
soliton D are linearly stable while soliton E (the antisymmetric counterpart
of the latter) is unstable.
Since in all cases considered $\tilde{\mu}_{0}=16.6$, it is clear from the
figure that the spectrum of this soliton exhibits a pair of purely
imaginary eigenfrequencies, indicating an exponential instability.
As for soliton F, even though the effective model (\ref{E1DE}) still predicts
correctly the (oscillatory) unstable nature of this soliton, it cannot give a
quantitative account of its 3D excitation spectrum which now contains a
quartet of complex frequencies not appearing in the corresponding
1D spectrum (which exhibits another quartet of complex frequencies, although
only a pair is visible in the field of view of the figure).
One expects these complex frequencies not appearing in the 1D spectrum to be
associated with unstable transverse excitations of the embedded dark soliton
(snake-like instability). To verify this and the previous conclusions we have
introduced an additive Gaussian white noise to produce a random small
perturbation in the stationary configuration of the different gap solitons and
have followed the ensuing time evolution by integrating numerically both the
effective 1D equation (\ref{E1DE}) and the 3D GPE (\ref{3DGPE}). The results
obtained are collected in Fig. \ref{Fig8}.
The upper left panel shows the dynamical evolution of soliton D in terms of a
density map where brighter regions correspond to higher densities. This image,
which has been obtained from the 1D effective equation (\ref{E1DE}), is
indistinguishable from the corresponding 3D result and confirms the stability
of the symmetric composite soliton D as predicted by the linear stability
analysis of Fig. \ref{Fig7}. The same holds true for the fundamental soliton B
(not shown in the figure). For later convenience, the upper right panel
displays the time evolution of the real part of the wavefunction of soliton D
(in arbitrary units) obtained from both the 1D (top) and the 3D models
(bottom). Since this soliton is stable, its dynamical evolution, apart from
small local fluctuations induced by the Gaussian perturbation, is given
essentially by a periodic amplitude modulation $\mathrm{\operatorname{Re}%
}(\phi(t))\simeq\phi(0)\cos(\mu t/\hbar)$ with period $T\simeq0.2$ ms, as
corresponds to a (quasi) stationary state. Thus stability implies, in
particular, that the relative phase between any pair of points in the global
amplitude envelope must remain (quasi) constant over time. Accordingly, even
though an analysis of the density evolution of the antisymmetric composite
soliton E (middle left panel in Fig. \ref{Fig8}) seems to indicate that this
soliton is stable (the same conclusions follow from both 1D and 3D results), a
more detailed study including also phase information reveals that this is not
true. Indeed, as can be seen from the middle right panel, in this case the
time evolution of the real part of the wavefunction no longer is a periodic
amplitude modulation. While the two major peaks are initially out of phase,
both the 1D effective model (top) and the 3D model (bottom) predict that they
become essentially in phase at $t\approx4.5-5$ ms, so that, this soliton turns
out to be unstable, in agreement with the linear stability analysis of Fig.
\ref{Fig7}. This example clearly reflects the importance of the spectrum of
elementary excitations in the analysis of the dynamical stability of these
nonlinear structures.
In the two lower panels we compare the evolution in time of the density of
soliton F obtained from Eq. (\ref{E1DE}) (left panel) with that obtained from
the 3D GPE (right panel). As is apparent, in this case the effective 1D model
predicts a longer soliton lifetime than the 3D model, in good agreement\ with
the linear stability analysis of Fig. \ref{Fig7}. The inset in the right panel
shows that the soliton decay occurs when the nodal plane of the embedded dark
soliton begins to undergo a considerable deformation, thus, indicating that
the decay is a consequence of a transverse modulational instability which
clearly cannot be accounted for with a 1D model.
One may wonder why the effective model (\ref{E1DE}) is able to properly
predict the stability properties of the gap solitons even though, in the
presence of the optical lattice (and because of the high frequency axial
motion it might induce), in general, it is unable to accurately reproduce
the detailed evolution in time. To address this question, we begin by noting
that the adiabatic (Born--Oppenheimer) approximation essentially consists in
assuming that the axial dynamics is so slow in comparison to the typical time
scale of the radial degrees of freedom that, at every instant $t$, the latter
can adjust to their equilibrium configuration compatible with the axial
configuration $n_{1}$ occurring at that instant. This is exactly what the
stationary radial equation (\ref{II-7}) reflects. In the presence of an
optical lattice, one thus reasonably expects this approximation to work better
in those situations where the time variable does not play a prominent role,
such as occurs for the stationary solutions of the time-dependent GPE, which
are completely determined from a (time-independent) eigenvalue problem.
Something similar can be expected for the linear stability properties of the
system. Indeed, such properties are completely characterized by the frequency
spectrum of the elementary excitations, which follow from the solution of an
eigenvalue problem (BdG equations) where time does not play any role.
Even though an instability manifests itself in the evolution in time, the very
existence of a certain unstable mode is a characteristic feature of the system
which depends only on the interplay among the various energies and interactions
contributing to the Hamiltonian. This explains why the effective model
(\ref{E1DE}) can predict the stability properties of the gap solitons even
though it is unable to properly account for their dynamical evolution.
\section{V. CONCLUSION}
Most theoretical studies of gap solitons realized in condensates loaded in 1D
optical lattices have been carried out in terms of the 1D GPE. This is
motivated, in part, by the analogy of the latter equation with the 1D NLSE
governing the evolution of similar structures in nonlinear optics and, in
part, by the fact that, as occurs with dark solitons, one would reasonably
expect these structures to be more stable in a quasi-1D regime, where their
potential decay would be inhibited by the strong radial confinement. However,
it has been demonstrated that robust, long-lived gap solitons exist in a 3D
regime where many higher-order radial modes are excited \cite{GS-3D}.
Moreover, matter-wave gap solitons are intrinsically 3D and in realistic
situations can hardly satisfy conditions (\ref{con3}), which are necessary for
the validity of the 1D GPE. In these circumstances, effective equations like
Eq. (\ref{E1DE}), which can take the 3D character of stationary
BECs into account by incorporating contributions from higher-order radial
modes, offer a clear advantage over the latter. However, the question remains
open whether the effective 1D equation (\ref{E1DE}) represents a reliable
alternative to a fully 3D treatment. In this work we have addressed this
question by analyzing the ability of the above effective equation to reproduce
the fundamental physical properties of stationary matter-wave gap solitons in
realistic conditions. Our results demonstrate that the predictions from the
effective model (\ref{E1DE}) are in good quantitative agreement with the exact
numerical results from the full 3D GPE. In particular, unlike the standard 1D
GPE (which fails in most cases of practical interest), the 1D effective model
(\ref{E1DE}) correctly predicts the distinctive trajectories characterizing
the different gap soliton families as well as the corresponding axial
wavefunctions along the entire band gaps. Our results also show that this
effective model can predict correctly the stability properties of the
different gap soliton families. This follows from both a linear stability
analysis in terms of the corresponding BdG equations and a representative
set of numerical computations monitoring the long time response of the
system to a sudden small random perturbation.
In particular, by numerically solving the BdG equations we have
analyzed the prediction the effective model makes for the spectrum of
elementary excitations, which proves to be essential for unambiguously
determining the stability properties of certain matter-wave gap solitons. Our
results show that, except in those cases where the soliton features a
nontrivial (unstable) transverse configuration, the model is able to correctly
predict the spectrum of complex eigenfrequencies responsible for the dynamical
stability of the system. These results thus indicate that the above effective
1D model can be a useful tool for the physical description of realistic
matter-wave gap solitons in 1D optical lattices.
As a natural extension of the present work, one might consider to quantitatively
investigate the effects of quantum fluctuations on the above stationary
matter-wave gap solitons. In the tight-binding regime ($s\gg1$), the tunneling
of atoms between adjacent lattice sites can be strongly inhibited. Under these
conditions, intersite atom-number fluctuations become significantly suppressed
while the corresponding phase fluctuations become enhanced, which can lead to a
reduced relative phase-coherence. While for fundamental gap solitons (those
located at a single lattice well) with a few hundreds atoms these quantum effects
are not expected to be very significant, they can be more important in the case
of composite in-phase or out-of-phase solitons. The problem of quantifying the
contribution of these quantum fluctuations can be addressed in terms of a
stochastic phase-space description based on the Truncated Wigner Approximation
(TWA) \cite{TWA}. In this respect, we note that since Eq. (\ref{E1DE}) properly
captures the Bogoliubov excitation spectrum of the system, it represents a
particularly convenient starting point for such an analysis. This will be the
subject of a future publication.
|
\section*{Results}
\subsection*{The adoption dynamics}
The spreading of the online service is determined by competing processes of adoption and termination as described by the evolution of the corresponding rates $R_a(t)$ and $R_t(t)$, which measure the fraction of all users that adopt or terminate the service in a given time window $\Delta t$ (Fig. \ref{fig:rates}.a). These simple rate functions already disclose interesting features of the adoption dynamics, since their overall growth signals continuously accelerating processes of adoption and termination. Yet the actual time evolution of spreading service is better characterized by the net adoption rate $R_n(t)=R_a(t)-R_t(t)$ (for an overview of all empirical quantities see Table \ref{table:pars}).
Opening a user account constitutes a single event in the decision-making process of an individual that is triggered by either spontaneous decisions, the influence of media or by the social environment \cite{Aral2009,Watts2002}. On the other hand, users may terminate their accounts for several reasons including vanishing demand or dissatisfaction, by switching to another product permanently, or by simply abandoning the service with a chance of re-adoption (e.g. due to loss of password or intention for lower monitoring). Some of these processes are observable by investigating the data. An example is shown in Fig. \ref{fig:nets}, where the contact network of Switzerland is further decomposed into sub-networks of adopted and terminated users. In the former some nodes appear disconnected, which indicates individuals that have adopted Skype prior to their friends. This so-called \textit{spontaneous} adoption, where individual factors and external media play a role, is a typical adoption pattern in the beginning of the process (Fig. \ref{fig:nets}.b). Alternatively, at the time of adoption many nodes have neighbours who are already existing users, a common pattern for later times (Fig. \ref{fig:nets}.e). This second scenario of \textit{peer-pressure} adoption indicates the possible influence of the social environment. In contrast, the termination network consists mostly of single nodes at all times (Fig. \ref{fig:nets}.c and f), meaning that these users, although they are surrounded by adopters, decide individually to terminate. This observation suggests a negligible effect of social influence on termination.
\begin{figure}
\centerline{\includegraphics[width=.8\textwidth]{Fig2.pdf}}
\caption{(Online version in colour.) Empirical rates and probabilities for Switzerland. (a) Thin curves denote empirical rates of adoption [$R_a(t)$], termination [$R_t(t)$], and net adoption [$R_n(t)$], while symbols are their corresponding binned values. A binned data point in $[2T,3T]$ has been removed due to systematic bias in $R_t(t)$ caused by a major software update during this period. A shaded (white) area indicates the training (predicted) period for the theoretical fit of our model, drawn as thick lines with the same colors as the empirical rates. (b) Probabilities of spontaneous [$p_a(t)$] and peer-pressure [$p_p(t)$] adoption per unit time. (c) Average conditional probability of adoption as a function of the fraction of adopting neighbors $n$, measured in the original data [$p(n)$, solid circles] and in the shuffled data corresponding to the null model [$p_{\mathrm{rand}}(n)$, open circles]. Inset shows the unbiased difference $\Delta p(n) = p(n) - p_{\mathrm{rand}}(n)$ (symbols) and a fitted linear function (continuous line). (d) Probabilities of overall termination [$p^-(t)$], and of spontaneous [$p_a^-(t)$] and peer-pressure [$p_p^-(t)$] termination per unit time. The inset depicts a zoom from time $2T$ onwards. $T$, $r$ and $c$ are arbitrary linear scaling constants, with time dimensions for $T$. Black lines in panels (b), (d) are fitted constants.}\label{fig:rates}
\end{figure}
\subsection*{Mechanisms of adoption}
An analysis of the evolving network structure around a given user can help us to detect whether an ego adopted or terminated the product before any of its neighbours did; or else followed the decisions previously made by a fraction of them. In this way we can label the performed action as either spontaneous or driven by peer pressure. To define the related measures we consider the underlying social network as static, meaning that its evolution requires a much larger temporal scale than the adoption process itself. This static structure is defined as the aggregated social network of Skype at the end of the recorded period, and provides a lower estimate for the total number of friends of each individual. Moreover, we assume that the maximum size of the static social network is the number $I$ of internet users in a given country at the end of the observation period (2011) \cite{Internet}, and thus define $I - N_a(t)$ as the population that has not yet adopted Skype at time $t$.
Under these assumptions, the probabilities per unit time that a user adopts either spontaneously or due to peer pressure are defined as,
\begin{equation}
\label{eq:empAdopt}
p_a(t)=\dfrac{\# ad(t+\Delta t|SF=0)}{I-N_a(t)}, \quad \text{and} \quad
p_p(t)=\dfrac{\# ad(t+\Delta t|SF\neq 0)}{I-N_a(t)},
\end{equation}
where $\# ad(t+\Delta t|SF=0)$ [$\# ad(t+\Delta t|SF\neq 0)$] is the number of users who adopt the service in a time window $\Delta t$, under the condition that their number of adopting neighbours at time $t$ is $SF=0$ ($SF \neq 0$). In a similar fashion, the probabilities per unit time that a user terminates the service either spontaneously or due to peer pressure are,
\begin{equation}
\label{eq:empTerm}
p_a^-(t) = \dfrac{\# tr(t+\Delta t,TF=0)}{N_a(t)}, \quad \text{and} \quad
p_p^-(t) = \dfrac{\# tr(t+\Delta t,TF\neq0)}{N_a(t)},
\end{equation}
where $TF$ stands for the number of neighbours of a user that have terminated usage up to time $t$ (for a discussion on the restrictions of these empirical quantities see Supplementary Section S2.1-S2.4).
The data shows that after an initial, transient period, the rate of spontaneous adoption $p_a(t)$ (Fig. \ref{fig:rates}.b) and the rate of termination $p^-(t)=p_a^-(t)+p_p^-(t)$ (Fig. \ref{fig:rates}.d) become constant apart from small fluctuations. The same holds separately for the rates of spontaneous [$p_a^-(t)$] and peer-pressure [$p_p^-(t)$] termination. The time invariance of these rates is an obvious assumption for most biological epidemics, which, however, has never been empirically shown before in the case of social contagion phenomena, despite its wide use \cite{Vespignani_review,Castellano2009}. Our results provide the first validation of this quite fundamental assumption used in the conventional modelling of social spreading processes, where probabilities analogous to the ones described here are treated like constants at the outset.
When the ego is not the first adopter among neighbours, the rate $p_p(t)$ of adoption via peer pressure is not constant but increases with time (Fig. \ref{fig:rates}.b). This is mainly due to social influence arising from the user's social circle. An appropriate way to quantify such effects is to measure the conditional probability $p(n)$ of adoption provided that a fraction $n$ of the ego's neighbours have adopted the product before as
\begin{equation}
\label{eq:pn}
p(n)=\dfrac{\# ad(n)}{N-\sum_{m=0}^{m<n}\# ad(m)}.
\end{equation}
Here the numerator counts the number of users with a fraction $n$ of adopter friends at the time of adoption, while the denominator is the number of people with a larger or equal fraction $m \geq n$, i.e. all individuals who had the chance to adopt Skype while having a fraction $n$ of adopter neighbours (for further details see Supplementary Section S2.3). We observe that the probability $p(n)$ is monotonically increasing (Fig. \ref{fig:rates}.c), an empirical finding in agreement with the assumptions of several threshold models for epidemic spreading and social dynamics \cite{Watts2002,Dodds2004,Klimek,Takaguchi2013}. However, since we cannot see the entire social network (only the part uncovered by the Skype graph), this probability is biased as $n \rightarrow 1$. To estimate such bias, we build a reference null model by shuffling the adoption times of all accounts and measuring the corresponding conditional probability $p_{\mathrm{rand}}(n)$ for this system. The shuffling procedure removes the effect of social influence but conserves the adoption rates and keeps the social structure unchanged. In other words, the reference probability is biased in the same way as the original measurement, but is not driven by social influence as all such correlations have been removed by the shuffling. Consequently the difference $\Delta p(n)= p(n) - p_{\mathrm{rand}}(n)$ quantifies the effect of social influence in the adoption process (inset of Fig. \ref{fig:rates}.c): $\Delta p(n)$ increases approximately in a linear fashion with the fraction of adopting neighbours. This observation is in agreement with previous studies where a similar scaling of social influence has been recognized through small scale experiments \cite{Centola2010}, data-driven observations \cite{Bakshy2012}, and modelling \cite{Dodds2004,Dodds2013}.
\input{Table1.tex}
\subsection*{The model process}
The analogy between epidemic spreading and social contagion has been widely used to model various societal diffusion processes \cite{Jackson2005,Hill,GoffmanNevill,DaleyKendall}. Here we take this approach to build a compartmental model based on the identified mechanisms in Skype usage, aimed at a generic description of the large-scale adoption dynamics of technological innovations. We depict individuals as agents in one of three non-overlapping states, susceptible ($S$), adopter ($A$) and removed ($R$), describing people who may adopt the product later, are users already, and will never use it again. In accordance with our observations, the behaviour of an agent can be characterized by four elementary processes: (a) \textit{Spontaneous adoption}, influenced by individual factors or external media independently of the social network. This is certainly the dominant mechanism for agents with no user neighbours at the time of adoption. (b) \textit{Peer-pressure adoption}, an intrinsic social effect implemented here by making use of the observed linear scaling of the probability $p(n)$. (c) \textit{Temporary termination}, describing the case in which agents stop usage with a chance of re-adoption. (d) \textit{Permanent termination}, when users abandon the service altogether. The flow $S \rightarrow A$ is regulated by processes (a) and (b), $A \rightarrow S$ by (c), and $A \rightarrow R$ by (d). Finally we assume that the underlying social network evolves with a much longer time scale than the ongoing adoption process, so that its structure may be considered static with fixed size.
For large systems, the modelled adoption process can be well characterized by a rate equation formalism using the heterogeneous mean-field approximation \cite{Barrat2008,Satorras2001} (see Appendix A). This approach takes agents with identical degree to be statistically equivalent and ignores fluctuations in their dynamical properties. Thus, assuming no degree-degree correlations in the network, the adoption dynamics is reduced to the following system of non-linear ordinary differential equations,
\begin{eqnarray}
\label{eq:uncorrODEs1}
da/dt & = & [p_a + p_{pk}(1 - p_a)a]s - [p_r + p_s(1 - p_r)]a, \\
\label{eq:uncorrODEs2}
ds/dt & = & -[p_a + p_{pk}(1 - p_a)a]s + p_s(1 - p_r)a, \\
\label{eq:uncorrODEs3}
dr/dt & = & p_ra,
\end{eqnarray}
where $s(t)$, $a(t)$, and $r(t)$ are the average probabilities that an agent is in state $S$, $A$ or $R$, respectively, and satisfy the normalization condition $s(t)+a(t)+r(t)=1$. The elementary mechanisms (a--d) are parametrized through the constant probabilities of spontaneous ($p_a$) and peer-pressure ($p_p$) adoption, and of temporary ($p_s$) and permanent ($p_r$) termination (for an overview of all model parameters see Table \ref{table:pars}). Under the above conditions the model does not depend on the degree distribution of the social network, since $p_p$ appears only in the weighted form $p_{pk} = p_p(\langle k \rangle - 1)/\langle k \rangle$, with $\langle k \rangle$ the average degree of the network. Moreover, for large $\langle k \rangle$ the model becomes independent of this quantity as $p_{pk} \sim p_p$. The system (\ref{eq:uncorrODEs1})--(\ref{eq:uncorrODEs3}) finally allows us to write the theoretical rates of adoption and termination as,
\begin{eqnarray}
\label{eq:uncorrRates1}
R_a(t) & = & [p_a+p_{pk}(1-p_a)a]s, \\
\label{eq:uncorrRates2}
R_t(t) & = & [p_r+p_s(1-p_r)]a,
\end{eqnarray}
that is, the gain and loss terms in Eq.~(\ref{eq:uncorrODEs1}) (detailed derivation in Supplementary Section S3).
In order to measure the effect of degree-degree correlations on the spreading dynamics, we perform data-driven simulations by evaluating the model process over the integrated Skype network of a country (Fig. \ref{fig:model}.a). While this empirical network retains its full topological complexity (in terms of real community structure, assortativity, etc.), we consider an model scale-free network \cite{barabasi1999} of the same size and average degree as a control case. We then run the model process over the empirical and control networks, and compare their corresponding rates of adoption and termination with the mean-field prediction of Eqs.~(\ref{eq:uncorrRates1}) and~(\ref{eq:uncorrRates2}). Since the average degree of the Skype network is not too small, deviations of the simulated rates from the theoretical values are not large, resulting in rates with the same qualitative behaviour. More interestingly, there is only a small discrepancy between the rates of the empirical and control networks. This suggests that topological correlations have a minor slowing-down effect on the spreading dynamics, and thus play a negligible role in the overall rates of the adoption process. Note that a similar independence of the population structure has been observed earlier in controlled experiments of networked public goods games \cite{Suri2011} and a spatial Prisoner's Dilemma \cite{gracia2012}. These observations validate retrospectively the theoretical considerations mentioned above, where we have assumed the background social network to be uncorrelated.
\subsection*{Validation and socio-economic correlations}
In order to validate our model process, we compare the theoretical rate functions of Eqs.~(\ref{eq:uncorrRates1}) and~(\ref{eq:uncorrRates2}) with the empirical data by estimating some of the model parameters. An estimate $\widetilde{\langle k \rangle}$ for the average degree of the social network can be obtained from the fully aggregated Skype network of a country, if we consider the ego-network of each user with international links included. The estimated rate of termination, $\widetilde{p^-}$, can be measured from the long time behaviour of the spreading process (Fig. \ref{fig:rates}.d) and then used to fix $p_s$ to the value $\widetilde{p_s}=(\widetilde{p^-}-p_r)/(1-p_r)$, where $p_r$ is a free parameter. While $p_a$ could also be measured directly (by counting adoption events where no user neighbours are present at the time), the observation time of the dataset is not sufficiently long to estimate a constant value $\widetilde{p_a}$ in all countries. Therefore we leave $p_a$, $p_p$ and $p_r$ as free quantities to be fitted (estimation of $p_a$ for selected countries in Supplementary Section S4).
\begin{figure}
\centerline{\includegraphics[width=.6\textwidth]{Fig3.pdf}}
\caption{(Online version in colour.) Data-driven simulations and validation of model predictions. (a) Rates of the model process over the integrated Skype network of Switzerland (dark symbols) and over an uncorrelated scale-free network (light symbols) with similar size and average degree. Parameter values for the numerical simulations are arbitrarily set to $p_a = 0.00004$, $p_p = 0.0024$, $p_s = 0.0004$ and $p_r = 0.0008$. Solid lines indicate the theoretical rates calculated from Eqs.~(\ref{eq:uncorrRates1}) and~(\ref{eq:uncorrRates2}). Similar scaling of the data-driven, simulated, and theoretical curves suggests a minor effect of structural correlations. (b), (c) Comparison between empirical and theoretical values of the rates of adoption and termination for 34 different countries. Symbols depict rates averaged over the last six months of the observation period, with their corresponding standard deviations as errorbars. In-symbol letters are country abbreviations, colors denote continental territories, $r$ is an arbitrary linear scaling constant, and the dashed line is a linear function with unit slope.}\label{fig:model}
\end{figure}
Overall, the model dynamics is characterized by $\{ p_a, p_p, p_r, \widetilde{p_s}, \widetilde{\langle k \rangle} \}$, a set of three free parameters and two estimated quantities. The free parameters are used to simultaneously fit the model rates on the binned empirical rates $R_a(t)$, $R_t(t)$ and $R_n(t)$, by means of a bounded non-linear least-squares method. To ascertain the predictive power of our model, we fit over a 5-year training period and look for predictions in the last 1.5 years (Fig. \ref{fig:rates}.a). Such prognosis can be quantified by comparing the average rates provided by the model with their corresponding empirical values during the final six months of observation. After repeating the calculations for 34 different countries (with diverse levels of technological development), the related values of the final empirical and modelled rates all collapse close to a line with unit slope (Fig. \ref{fig:model}.b and c), thus validating our model for the studied adoption process.
\begin{figure}
\centerline{\includegraphics[width=.8\textwidth]{Fig4.pdf}}
\caption{(Online version in colour.) Measures of innovation spreading vs. socioeconomic development in a country. (a) Predicted inverse speed of adoption spreading $\tau$ as a function of GDP per capita \cite{GDP} for various countries. (b) Theoretical inverse termination probability $1 / \widetilde{p^-}$ as a function of the empirical average account lifetime $\langle t_l \rangle$. (c) $\langle t_l \rangle$ as a function of a press liberty measure \cite{Liberty} (large scores imply weak liberties). $\bar{T}$ and $\bar{c}$ are arbitrary linear scaling constants with time dimensions for $\bar{T}$.}\label{fig:corrs}
\end{figure}
Our model may also be used to disclose relevant differences between the adoption dynamics of countries at various levels of societal and economical development. One characteristic indicator is the inverse speed of innovation diffusion, defined as the time $\tau$ when the theoretical $R_n(t)$ is maximal (see Supplementary Section S3.4). If we relate $\tau$ with one of the standard measures of economical development, GDP per capita \cite{GDP}, large differences emerge between countries (Fig. \ref{fig:corrs}.a). Specifically, the larger the GDP of a country, the faster the adoption process is in its society. Another way to characterize the adoption dynamics is through the average account lifetime, $\langle t_{l}\rangle=\langle t_t-t_a \rangle$, where $t_a$ and $t_t$ are the corresponding registration and termination times. We relate this empirical measure to its theoretical analogue, the inverse probability of termination $1/\widetilde{p^-}$ obtained from the fitted model process (Fig. \ref{fig:corrs}.b). Their correlation indicates that our model captures this dynamical property correctly. Moreover, the typical duration of user engagement uncovers clusters of countries at different levels of socio-economic development. This can be better understood by linking $\langle t_{l}\rangle$ with general civil liberty measures \cite{Liberty} (Fig. \ref{fig:corrs}.c). We observe that the weaker the press liberty is in a country, the shorter time online accounts are used there (other liberty measures in Supplementary Section S4.3). Such observations indicate a quantifiable dependence between the dynamics of innovation spreading and the socio-economic status of a country.
\section*{Discussion}
Our analysis of one of the largest online communication services worldwide aimed at clarifying several long-standing questions about the spreading mechanisms of novel technologies. We have shown that innovation diffusion can be interpreted as a competition between service adoption and termination; a process characterised, after a transition time, by constant rates and by a linearly increasing influence of user neighbours on service adoption. In addition we have integrated the identified mechanisms into a minimal modelling framework that provides accurate medium-term predictions for the spreading of an online service.
It should be pointed out that the present study has some limitations. First, the complete structure of society cannot be mapped by using online interactions only, as observations taken from any online social network underestimate the real number of contact peers of an ego. This incompleteness allows us only to estimate effective degrees and adoption thresholds. Second, the observation of correlated adoption does not necessarily imply the presence of actual social influence, only its possibility; even more so since other mechanisms like homophily cannot be synthesized from this dataset. Despite these limitations, the presented results provide strong evidence of key mechanisms driving the complex contagion of online technologies, up to a level of detail and scale that has not been possible before.
These results may help fill an enduring gap between the theoretical understanding and the empirical observation of social contagion phenomena, validating several earlier studies based on similar assumptions, like constant adoption rates and the effect of social influence. In addition, we have shown how the adoption of novel technologies is related to the societal and economical development of a country. Beyond the clear advantage of these observations for the design of marketing and business plans, they also provide further insight into the differences in the development of modern online societies.
\vspace{.1in}
{\bf Acknowledgments:} The authors gratefully acknowledge the support of M. Dumas and A. Saabas from STACC and Microsoft/Skype Labs, and from the ICTeCollective EU FP7 project. M.K. thanks R. Kikas for the data preparation and P. Gon\c{c}alves for useful comments. G.I. acknowledges the Academy of Finland for funding. J.K. thanks FiDiPro (TEKES) and the DATASIM EU FP7 project for support, and S. Fortunato for discussions. This research was partly funded by Microsoft/Skype Labs.
\begin{appendices}
\section{Model description}
For a static social network $G$ with degree distribution $\rho_k$, the probability that individual $i$ becomes a user is $p^+_i = p_a + p_p(1 - p_a)n_i$ with $n_i = N_i / k_i$. Here $N_i$ is the number of neighbours of $i$ that have already adopted the product and $k_i$ its degree. Furthermore, the probability that $i$ stops being a user is $p^-_i = p_r + p_s(1 - p_r)$. In the thermodynamic limit we assume that all agents with the same degree are statistically equivalent, allowing us to group individuals and write rate equations for each degree class $k$. We denote by $s_k$, $a_k$ and $r_k$ the average probabilities that a randomly chosen agent with degree $k$ is susceptible, adopter and removed, respectively. A first-order moment closure method leads to the rate equation $da_k / dt = \langle p^+_i \rangle s_k - \langle p^-_i \rangle a_k$. In other words, the average probability that an adopting agent becomes either removed or susceptible is $\langle p^-_i \rangle a_k = [p_r + p_s(1 - p_r)]a_k$, while the average probability that a susceptible individual adopts the product is $\langle p^+_i \rangle s_k = [p_a + p_p(1 - p_a)n_a]s_k$ with $n_a = \langle n_i \rangle$. This approximation ignores higher moments of the dynamical quantities $s_k$, $a_k$ and $r_k$, as well as any correlations between them. In the presence of degree-degree correlations in $G$ we have $n_a = \sum_{k'} (k' - 1) \rho_{k', k} a_{k'} / k'$, where $\rho_{k', k}$ is the conditional probability that an edge departing from an agent with degree $k$ arrives at an agent with degree $k'$. Similar rate equations can be written for $s_k$ and $r_k$, leading to a system of non-linear ordinary differential equations that determines adoption at the degree class level (for further details see Supplementary Section S3).
\end{appendices}
|
\section{Introduction}
\label{sect:intro}
Blazars are the most extreme class of AGNs exhibiting very violent high energy phenomena. The subset of blazars can be grouped into two very different categories: BL Lacertae-type objects (BL Lacs) and flat spectrum radio quasars (FSRQs). They have been observed at all wavelengths from radio through very-high energy (VHE) gamma-rays, and exhibited rapid variability at all wavelengths on various time scales, from years and a few months to even shorter than an hour in some cases. The emission from blazars is thought to be highly beamed through the relativistic jet. Flux variation study is considered to be a powerful tool for understanding the structure and emission mechanism of AGNs. Periodical variabilities on a wide range of timescales have been reported by some investigations, e.g. \citep{fan00,xie08,king13}. The most notable case is that of OJ 287, which has an over 120-year-long light curve and a convincing 11-12 year periodicity \citep{kidger92,valtonen06}. However, there is still some debate over whether blazar variability is periodic.
Multi-wavelength observations of blazars have been performed for many years. They can give the general shape of the energy spectrum (i.e. flux versus frequency) which follows a power law proximately. The energy spectrum can provide insight into the nature of the emission process. The low energy component between radio and optical even to X-ray frequencies, is mainly attributed to synchrotron emission from non-thermal electrons in a relativistic jet. The high energy emission between X-rays and TeV $\gamma$-rays, may be due either to the Compton up-scattering of low energy radiation by the synchrotron-emitting electrons \citep{botter07} or hadronic processes initiated by relativistic protons co-accelerated with the electrons \citep{mucke03}. Measurements of blazar spectral variability are important tools in constraining physical models. In the optical domain, the color index is often used to represent spectrum index, and its variations usually accompanies flux variabilities with different behaviours, such as "bluer when brighter" or "redder when brighter" trend.
The high-frequency peaked BL Lac PKS 2155-304 with a redshift of z = 0.116 is one of the brightest BL Lacs in the X-ray and EUV bands \citep{giommi98}. It was classified as a TeV blazar by the detection of VHE gamma rays by the Durham MK 6 telescopes \citep{chadwick99} , and then was confirmed by the H.E.S.S. collaboration with a high significant of 45$\sigma$ at energies greater than 160 GeV \citep{aharonian05}. This source has been observed on diverse timescales over a wide range of frequencies from radio to VHE gamma-rays, and shown rapid and strong variability \citep{dominici06,dolcini07, aharonian07,abramowski10, kastendieck11,abramowski12, aleksic12, barkov12}.
The long-term optical variability of PKS 2155-304 has been studied by some authors \citep{fan00, zhang96, kastendieck11}. \cite{fan00} investigated the periodic variations in the long-term optical light curves, found two possible periodicities of 3.7 years and 7.1 years. The short-term color varies complicatedly, and shows different behaviours at different epoches, e.g. \citep{carini92,pesce97,xie99,dolcini07,bonning12}.
This research is aimed to investigate the periodic variations and long-term color behaviour with the historical optical light curves of PKS 2155-304. The paper is organized as follows: In Section 2, the optical data are compiled and the historical light curve is constructed; In Section 3, the technique used for searching periodicity are described and the periodic variations are investigated; In Section 4, the long-term color behaviour is studied; and then the discussion and conclusions are given in Section 5.
\section{Optical data}
All available historical archival data of PKS 2155-304 have been collected from the following literatures:
\cite{griffiths79}, \cite{miller83}, \cite{brindle86}, \cite{hamuy87}, \cite{treves89}, \cite{mead90},
\cite{smith91}, \cite{carini92}, \cite{smith92}, \cite{jannuzi93}, \cite{urry93}, \cite{courvoisier95},
\cite{xie96}, \cite{heidt97}, \cite{pesce97}, \cite{bai98}, \cite{xie99}, \cite{bertone00},
\cite{tommasi01}, \cite{xie01}, \cite{gupta02}, \cite{dolcini07}, \cite{osterman07}, \cite{kastendieck11}. Up-to-date SMART optical data \citep{bonning12} have also been collected. In sum, 179, 759, 1382, 8674 and 590 data points in $U$, $B$, $V$, $R$ and $I$ bands are compiled, respectively.
The observations cover the time duration from 1979 to 2013, and the time interval is about 35 yr.
The mean magnitudes in $U$, $B$, $V$, $R$ and $I$ bands are 12.52 $\pm$ 0.39, 13.54 $\pm$ 0.46, 13.04 $\pm$ 0.42, 12.78 $\pm$ 0.40, 12.23 $\pm$ 0.28, respectively. The source shows violent activity. The amplitudes of the observed variability are $\Delta U$ = 1.50 mag, $\Delta B$ = 2.23 mag, $\Delta V$ = 2.57mag, $\Delta R$ = 2.40 mag, $\Delta I$ = 2.34 mag, respectively. The variations in different passbands show the similar trend. The light curves are well correlated with each other. The $R$ band light curves are displayed in two panels of Figure 1, so as to exhibit more details.
\begin{figure}
\centering
\includegraphics[width=10.5cm, angle=0]{ms1740fig1.eps}
\caption{The historical optical $R$-band light curve of PKS 2155-304. The solid line and the short dash line are the folded light curves with a 317-day period using all the data and those after MJD 53400, respectively.}
\label{lightcurve}
\end{figure}
It is clear that there are more intensive observations after MJD 52500. In Figure 1, one can see that the source exhibited oscillations with $\sim$0.7 mag throughout the duration of MJD 52500 - 54100, superposed on a general and slow brightening trend with a total magnitude of $\sim$2. Then the source began to fade slowly till MJD 55200 by a total of $\sim$2 magnitude, accompanied by some small flickering with $\sim$0.7 mag. Subsequently, a large and sharp outburst appeared from MJD 55200 to 55300, after that the source brightness decreased with some small amplitude fluctuations. It is obvious that the short term variations are superposed on the long-term and large variations.
\section{Periodicity analysis}
To search for periodic variations in the long-term light curve, epoch-folding technique has been employed. This technique was described by \cite{leahy83} and extended by \cite{davies90,davies91}. It is less affected by harmonics, and more effective for nonsinusoidal modulations than the Fourier transform technique which is most sensitive to a periodic component of sinusoidal form in shape. This method is also well suited for sparse and unevenly sampled data set. The $N$ data points of a time series are folded on a trial period and binned by phase. For the $i$th of $M$ phase bins, the mean $x_{i}$ and sample variance $\sigma_ {i}^{2}$ are computed, as is the overall mean, $<x>$. Then the $Q^{2}$ statistic is computed,
\begin{equation}
Q^{2}=\sum_{i=1}^{M}\frac{(x_{i}-<x>)^{2}}{\sigma_{i}^{2}},
\end{equation}
which is distributed similarly to the $\chi^{2}$ statistic. \cite{davies90} pointed out that the test statistic is valid only for large sample sizes. He proposed an improved method for detecting periodicities based on the $L$-statistic, giving a greater sensitivity when the number of data points is limited. The $L$-statistic is defined as,
\begin{equation}
L=\frac{(N-M)Q^{2}}{(M-1)(N-1)-Q^2},
\end{equation}
which obeys an $F$ distribution with $M-1$ and $N-1$ degrees of freedom. The large value of $L$ means that the corresponding trial period may be a true period in the light curve. It can be tested by calculating the false-alarm probability (the confidence with which one can reject the null hypothesis, i.e. no periodic component).
The $R$-band observations are used to search for periodicity because there are more data available in the $R$-band than others. For each trial period, the data are folded into 20 phase bins, and then the $L$-statistic are calculated for a range of trial periods from 1 d to 1400 d in steps of 1 d. The $L$ is plotted as a function of the trial period in the lower panel of Figure~\ref{period}. The strong peak at 317 $\pm$ 12 d (i.e. 0.87 yr), is clearly visible (the error corresponds to the half of the full width at half maximum (FWHM)), which means the periodicity $P$=317 d. The integer multiples of this period are also detected strongly at 631 d, 954 d and 1263 d which correspond to 2$P$, 3$P$ and 4$P$, respectively.
\begin{figure}
\centering
\includegraphics[width=10.5cm, angle=0]{ms1740fig2.eps}
\caption{Upper: $DCF$ as a function of time delay, $\tau$; middle: $V^{2}_{m}$ as a function of trial periods; lower: $L$-statistic as a function of trial periods for the optical R-band light curve. The vertical dotted lines are drawn to guide the eyes.}
\label{period}
\end{figure}
To confirm the periodicity of 317 d, Jurkevich method \citep{jurkevich71} and discrete correlation function (DCF) method \citep{edelson88} are also applied to search for periodic signals in the light curve. The Jurkevich method adopted the folding technique, while it is based on the expected mean square deviation. The light curve is folded according to test periods and then all data are grouped to $20$ phase bins. The sum variance $V^{2}_{m}$ = $\sum_{i=1}^{20} $$V_{i}^{2}$ of all phase are computed and shown in the middle panel of Figure~\ref{period}. The minimum of $V^{2}_{m}$ suggests that the trial period may be true one. Obvious dips locate at 318 d, 630 d, 951 d and 1270 d, which represents the principal period, $P$, and its integer multiples, respectively. They are consistent well with those detected by epoch-folding method.
The DCF method is an useful tools to investigate not only correlations between different light curves but also periodic component in a light curve. For two time series, $a$ and $b$, $DCF(\tau)$ is defined as the mean of $(a_{i}-\overline{a}) (b_{i}-\overline{b}) /\sigma_{a}\sigma_{b}$ for all pairs with $\tau - \Delta \tau/2\leq t_{j}-t_{i}<\tau + \Delta\tau/2$. When $a=b$, the position of $DCF$ peak gives the periodicity information in the light curve. The $DCF$ for the $R$ band is computed and plotted as a function of $\tau$ (the upper panel of Figure~\ref{period}). The peak at $P=280$ d can be seen obviously. It is a little different to the period of 317 d derived by the epoch-folding technique. The subsequent peaks appear at 630 d, 920 d and 1270 d, which are near to the corresponding values detected by the other two methods. The last two peaks are very low, which means that they are modulated by a longer-term variation trend.
\section{Color behavior}
The behavior of colors provides a clue to understand the mechanism of time variations in blazars.
Since the blazars varies rapidly, the color index should be ideally evaluated using simultaneous observations in the different bands. However, there are few exactly simultaneous data. The color indices with quasi-simultaneous data have been calculated. The mean values of color indices are listed in Column 2 of Table~\ref{regression}. The correlation between the color and magnitude is investigated. Figure \ref{color} displays the color indices versus magnitudes. Linear regression analysis is performed, and the results are listed in Table~\ref{regression}. Columns 3 to 6 list the slope, the correlation coefficient $r$, chance probability $Prob.$ and number $N$, respectively.
\begin{table}
\bc
\caption[]{The results of linear regression analysis between color index and magnitude.
\label{regression}}
\setlength{\tabcolsep}{10pt}
\small
\begin{tabular}{cccccc}
\hline\noalign{\smallskip}
Color & mean & slope & $r$ & $Prob.$ & $N$ \\
\hline\noalign{\smallskip}
$B-R$ & 0.67 $\pm$ 0.09 & 0.066 $\pm$ 0.008 & 0.33 & $<10^{15}$ & 635 \\
$B-V$ & 0.32 $\pm$ 0.08 & 0.056 $\pm$ 0.007 & 0.30 & 3.56$\times 10^{-15}$ & 678 \\
$V-R$ & 0.33 $\pm$ 0.06 & 0.029 $\pm$ 0.004 & 0.23 & 2.51$\times 10^{-11}$ & 837 \\
\noalign{\smallskip}\hline
\end{tabular}
\ec
\end{table}
According to the results listed in Table~\ref{regression}, one can see that data points in Figure~\ref{color} are fitted well by straight lines. The small correlation coefficients suggest that there is a weak correlation between color index and magnitude.
There is a general indication that the object becomes bluer when it is brighter.
\begin{figure}
\centering
\includegraphics[width=10.5cm, angle=0]{ms1740fig3.eps}
\caption{Color indices versus magnitude of PKS 2155-304. The solid lines represent the best linear fitting.}
\label{color}
\end{figure}
\section{Discussion and conclusions}
Three methods are applied to search for periodic component in the $R$-band light curve. The results are consistent well with each other. A 317-day (i.e. 0.87 yr) periodic oscillation can be seen clearly. For the periodicity of 317 d, the corresponding $L$ is 519. This means that the chance probability of obtaining such a large value is extremely small. According to the $F$-distribution, the chance probability with the null hypothesis is less than $10^{-20}$. So, 317 day is suggested to be a true periodicity. In addition, this value can be confirmed by Jurkevich method. \cite{kidger92} provided a fraction $f=(1-V^{2}_{m})/V^{2}_{m}$ to assess the significance of the periodicity derived by Jurkevich method. A value of $f\geq0.5$ implies a strong periodicity in the light curve. In this analysis, $f=1.2$ suggests 317-day periodicity is true.
In Figure~\ref{lightcurve}, from MJD 53500 forwards, one can see five obvious peaks well separated by nearly 317-day intervals. The data shows the average separation is 316 days among these five peaks which are spaced by 321 day, 314 day, 325 day and 304 day, respectively. Going backwards and forwards from these five peaks, the periodicity seems to disappear. But after MJD 55500, the periodicity appears again with 328-day and 315-day intervals. Sometimes, the positions of the peaks are difficult to identified exactly. So, the folded light curve is calculated with the period of 317 days (Figure~\ref{phase-period}), and superimposed on the light curve in Figure~\ref{lightcurve} (the general shape of the whole light curve is taken into consideration). One can see that, except the peaks near MJD 53250 and MJD 55060, the folded light curves are consistent well with most of the peaks and the dips in the light curve, but nevertheless, the last three peaks seem to not be in the same periodic sequence with the previous ones. So 317 day is not a strict and persistent periodicity in the light curve. The variations with time scale longer than 317 days can also be seen from Figure~\ref{lightcurve}. This tells us that in the light curve there may be a longer periodic component which is difficult to be dug out at present.
\begin{figure}
\centering
\includegraphics[width=10.5cm, angle=0]{ms1740fig4.eps}
\caption{The folded light curves with the period of 317 days. (a): All the data are employed; (b): The data after MJD 53400 are employed. The curves seem to be in compliance with a sinusoidal wave.}
\label{phase-period}
\end{figure}
For PKS 2155-304, a 5-day roughly continuous light curve in November 1991 appeared to have a quasi-periodicity of 0.7 day \citep{urry93}. However, this periodicity was not confirmed by a more rigorous analysis, although more data covering the whole November 1991 were adopted \citep{edelson95}. Using the historical data from 1977 to 1994, \cite{fan00} found two possible periodicities of 4.16 years and 7.0 years in the $V$ band light curve. The two possible periodicities are respectively 5 and 8 times the periodicity (0.87 year) found in this work. They also pointed out the hints of period less than 4.0 years which could not confirmed at that time due to the limitation of data sample. \cite{kastendieck11} characterised the optical behaviour of PKS 2155-304 with a long-term light curve, however, they found no clear evidence for periodic behavior on any timescales.
Several models have been developed to explain the periodic variations, for example supermassive binary black hole model\citep{valtaoja00,xie02} and helical jet model (see \cite{rani09} and references therein). The quasi-periodic variations seem to suggest that there is a supermassive binary black hole (SMBBH) system in the center of a source.
The periodic brightness fading is caused by the eclipse of the system. In the case of PKS 2155-304, the emission is dominated by synchrotron emission and inverse Compton emission which both arise from relativistic jets. The quasi periodic variations are mostly caused by blobs propagating in a helical jet. The effect is similar to that of a jet whose direction changes. When the viewing angle is small, the source brighter, and vice versa. In this case, one would expect to see a rather exact variability pattern for some periods until the blob vanishes, and then a repetition with a new blob. From Figure~\ref{lightcurve}, the light curve seems to have a possible 5-peak stretch of almost periodic variations with a periodicity of 317 days, a break in the pattern, and then a 3-peak repetition of the 317-day periodicity, again followed by a break. These phenomena are consistent well with the expectation.
In general, there are five kinds of color behaviors:
(1) bluer-when-brighter (BWB) in the whole data sets. It is a well-observed feature in blazars especially in BL Lac object \citep{gu06,rani10,zhang10,ikejiri11}. This trend is generated by a variation component with a constant and relatively blue color and an underlying red component \citep{ikejiri11};
(2) redder-when-brighter (RWB) in the whole data sets. Most of FSRQs follow this redder-when-brighter tendency, which suggests the presence of a steady blue accretion disk component underlying the more variable jet emission \citep{rani10,bonning12};
(3) cycles or loop-like pattern, e.g. S5 2007+777 and 3C 371 \citep{xilouris06}, S5 0716+714 \citep{wu07}, and OJ 287 \citep{bonning12}, which may be caused by different amplitude and time delay in different spectral bands;
(4) redder-when-brighter at the low state while bluer-when-brighter at the high state (RWB to BWB), e.g. AO 0235+164 \citep{bonning12}, PKS 0537-441 \citep{zhang13};
(5) stable when brighter (SWB) or no correlation with brightness in the whole data sets \citep{ikejiri11,gu11}.
Since 1970s, color of PKS 2155-304 has been investigated by several authors. The color behavior showed very different and complex tendencies on different time scales and during different periods. On short time scales of days to months, during 1990 October and December, PKS 2155-304 showed a tendency to be bluer when the object was brighter (BWB)\citep{smith91}, but the tendency was not observed in 1990 November, the optical spectral index remained relatively constant even though the object brightened by nearly one magnitude (SWB) \citep{smith92}. Observations during 1991 November showed a constant spectral slope in the $U-I$ domain (SWB) \citep{courvoisier95}. During the period 1994, the average colors remained relatively constant, with no correlation with brightness (SWB) and slightly BWB \citep{pesce97}. During the campaign in 1995, there was clear evidence of hardening when the source got brighter (BWB) \citep{paltani97}. During 1996 August-October, bluer-when-brighter trend (BWB ) was observed by \cite{xie99}. From May to December 2005, no apparent correlation between spectral index and brightness (SWB) was found by \cite{dolcini07}, but the source exhibited a rather soft spectral shape during its high state (RWB).
On long time scales of years, this source during 1979-1982 became redder as it brightened (RWB) \citep{miller83}. From 1983 to 1985, it was found that the
higher state was harder than the lower one (BWB) \citep{treves89}. \cite{zhang96} collected the pre-1994 observations, and found no correlation between brightness and colors (SWB). Between 2001 and 2003, the optical colors showed a BWB phenomena but opposite ones in the infrared domain \citep{dominici06}. \cite{ikejiri11} found this source exhibited BWB trend during 2008 and 2010. The observations from 2008 to 2010, the overall trend over several years revealed no strong correlation between color and brightness (SWB) \citep{bonning12}. The data from April 2005 to June 2012 observed by REM telescope showed the color did not varied with the brightness (SWB) \citep{sandrinelli14}.
It is clear that PKS 0537-441 exhibited BWB or SWB trends most of time. In our color-magnitude analysis, the data covering 35 years is used. On the longest time scale till now, the source shows a clear bluer-when-brighter (BWB) trend which means the spectrum hardens when the source brightens. \cite{rani10} suggested that BWB and RWB both can be accommodated within shock-in-jet models.
In conclusion, the long-term possible periodic variations have been investigated, a 317-day (0.87 yr) periodic component in the light curve has been detected, and it is more convincing. On the whole, the source varies with a 0.87 yr periodicity superposed on a long-term slower trend. PKS 2155-304 shows complex color behaviour on different time scales. This analysis suggest a clear bluer-when-brighter chromatism on the long-term time scale.
\normalem
\begin{acknowledgements}
We thank the referee for great helps. This work is supported by the National Natural Science Foundation of China (11273008).
\end{acknowledgements}
\bibliographystyle{raa}
|
\section{Introduction}
\subsection{The problem}
In multivariate extensions of GARCH models, modelers are faced with the problem of correlations (between asset returns, in most applications).
The simplest idea is to assume that these correlations are constant in time, and constitute only an additional matrix of parameters.
This has provided the class of Constant Conditional Correlations models (CCC), first introduced by Bollerslev (1990). Since CCC models can be seen as the components of first-order Markov processes, once such models are rewritten in an extended vector space, it is relatively easy to prove the existence of strictly stationary and explicit solutions, even if the latter are analytically complex: see classical textbooks,
for instance Francq and Zako\"{\i}an (2010).
\mds
It rapidly became apparent that the assumption of constant correlations is too strong. It does not correspond to economic intuition or many
empirical features: see the recent paper by Otranto and Bauwens (2013) and the numerous references therein, for instance. Therefore,
Engle (2002) proposed extending CCC specifications by adding particular dynamics on the (conditional) correlation matrices of returns, denoted here by $(R_t)$. To ensure modelers are dealing with true correlation matrices, he introduced a non-linear transform: there exists a sequence of
variance-covariance matrices $(Q_t)$ such that $R_t =diag(Q_t)^{-1/2} Q_t diag(Q_t)^{-1/2}$, and $(Q_t)$-dynamics are specified instead of $(R_t)$-dynamics directly, contrary to other authors (Tse and Tsui, 2002 or Pelletier, 2006, for instance). This non-linear transform ensures that $R_t$ is always a correlation matrix, i.e. positive semidefinite with ones on its main diagonal. Nonetheless, it considerably complicates the work of stating DCC model stationarity conditions. Indeed, analytically tractable solutions to such processes no longer exist. This explains why the existence and uniqueness of DCC model stationarity solutions have not yet been established in the literature, nor have the finiteness of their
moments. Particularly, this implies that theoretically sound statistical inference procedures do not yet exist, as noted in Caporin and McAleer (2013).
\mds
Despite their theoretical shortcomings, DCC models have been used intensively among academics and practitioners. Besides numerous applied works, several extensions of the baseline DCC representation have been proposed in the literature: inclusion of asymmetries (Cappiello, Engle, and Sheppard, 2006), volatility thresholds (Kasch and Caporin, 2013), macro-variables (Otranto and Bauwens, 2013), univariate switching regime probabilities (Pelletier, 2006, Billio and Caporin, 2005, Fermanian and Malongo, 2013), among others. Other authors have revisited the DCC parameterization itself: Billio, Caporin, and Gobbo (2006), Franses and Hafner (2009), etc. Therefore, there is an urgent need for new theoretical
results concerning the seminal DCC model itself.
\mds
Usually in econometrics, proving the existence of stationary solutions is the first step towards developing a full asymptotic theory (consistency/asymptotic normality of QML estimates typically, as in Comte and Lieberman (2003) in the case of multivariate GARCH models), because laws of large numbers and some CLTs are easily obtained in this case. In the GARCH literature, this essential task has been fulfilled by Bougerol and Picard (1992) for univariate GARCH models, by Ling and McAleer (2003) for multivariate ARMA-GARCH models, and by Boussama et al. (2011) for BEKK models. In the case of DCC models, a keystone is missing: a theory for inference has been proposed by Engle and Sheppard (2001), but their two stage estimation procedure is contingent on the underlying DCC process being strictly stationary and ergodic (see their Assumption A.2). The goal of this paper is to fill this gap.
\mds
After introducing some notations, we define DCC models at the beginning of Section~\ref{DCCrewriten}. They will be rewritten as
``almost linear'' Markov chains in Subsection~\ref{DCCMarkov}. The existence of strong and weak stationary solutions is stated in
Subsection~\ref{MainRes}. Subsection~\ref{Unicity} exhibits sufficient conditions to get their uniqueness. We discuss the tightness of our technical conditions from a qualitative standpoint in Section~\ref{My_discussion}. Proof of the propositions and theorems are detailed in the appendices.
\mds
\subsection{Notations}
\label{notations}
Consider an $(n,m)$ matrix $M=[m_{ij}]_{1\leq i
\leq n, 1\leq j \leq m}$.
\begin{itemize}
\item $M\geq 0$ (resp. $M>0$) means that all elements of $M$ are non-negative (resp. strictly positive), and $|M|=[|m_{ij}|]_{1\leq i
\leq n, 1\leq j \leq m}$.
\item If $n=m$, let the diagonal matrix $diag(M)=[m_{ij}\1 (i=j)]_{1\leq i
\leq m, 1\leq j \leq m}$ and the vector $Vecd(M)=[m_{ii}]_{1\leq i \leq m}$ in $\RR^m$.
\item If $n=m$ and $M$ is symmetric, $Vech(M)$ denotes the $m(m+1)/2=:m^*$ column vector whose components are read from $M$ column-wise and without redundancy.
To be formal, $Vech(M)=[\tilde m_{k}]_{1\leq k \leq
m^*}$, where $\tilde m_{k}=m_{ij}$ for the unique couple of
indices $(i,j)$ in $\{1,\ldots,m\}^2$, $i\geq j$ such that $ [m
+ (m-1) + \ldots + (m-j+2)]^+ + (i-j+1) =k$. This defines
a one-to-one mapping $\phi$ between the indices
$k\in\{1,\ldots,m^*\}$ and the pairs $(i,j)$, $i\geq j$, $1\leq
i,j \leq m$, i.e. $(i,j)=(\phi_1(k),\phi_2(k))=:\phi(k)$.
\item $\otimes$ denotes the usual Kronecker product, and
$M^{\otimes p} = M\otimes \ldots \otimes M$ ($p$ times). $\odot$ denotes the element-by-element product. If $v$ is a
vector in $\RR^n$, then $v\odot M = [v_i m_{ij}]_{1\leq i \leq
n,1\leq j\leq m}$.
\item We will consider several matrix norms, particularly
$$\|M\|_{\max} = \max_{1\leq i\leq n,1\leq j\leq m}|m_{ij}|,$$
and the spectral norm, defined for any squared matrix by
$$ \| M \|_s = \sup\{\sqrt{\lambda}\;| \; \lambda \;\text{is an eigenvalue of}\; M'M\}=\sup_\xx \frac{\|M\xx\|_2}{\| \xx \|_2}\cdot$$
Besides, we will consider any norm $\Nc$ for vectors, that is not the Euclidian norm $\|\cdot\|_2$. Then, we can define the norm $\|\cdot\|_{\Nc}$ for matrices by setting
$ \| M \|_{\Nc} = \sup_\xx \Nc(M\xx)/\Nc(\xx).$ Note that $\|M\|_{\infty}= \max_{i} \sum_{j} |m_{ij}|$ when $\Nc(\xx)=\|\xx\|_{\infty}=\max_i |x_i|$.
\item $\rho(M)$ denotes the spectral radius of the squared matrix
$M$, i.e. the largest of the modulus of $M$'s eigenvalues. If
$M$ is positive semi-definite, then $\rho(M)= \|M\|_s$ and its smallest eigenvalue is
denoted by $\lambda_1(M)$.
\item For any column vector $z_t\in \RR^m$, we denote
$z_t=(z_{1,t},\ldots,z_{m,t})'$ and
$\vec{z}_t:=(z_{1,t}^2,\ldots,z_{m,t}^2)'$.
\item $e$ denotes a vector of ones, the dimension of which will be
implicit. $0_m$ (resp. $I_m$) denotes the $m\times m$ matrix of zeros (resp. identity
matrix). When the dimension of an identity matrix is not
specified, it will be denoted by $Id$.
\item If $M$ depends on $\xx\in A$, then $\sup_{\xx \in A} M(\xx)$ is the matrix
$[\sup_{\xx\in A} m_{ij}(\xx)]$.
\end{itemize}
\mds
\section{Dynamic Conditional Correlation models}
\label{DCCrewriten}
\subsection{The classical DCC specification}
Let us reiterate here the standard DCC model, as introduced in Engle (2002). Consider a stochastic process $(y_t)_{t\in\ZZ}$ in $\RR^m$, typically a vector of $m$ asset returns. The sigma field generated by the past information of this process until up to and including time $t-1$ is denoted by $\Ic_{t-1}$.
Modeling the expected returns of financial series is a problem per se,
that has generated a huge amount of literature. In this paper, our
focus will be on the dynamics of the conditional variance-covariance
of $y_t$ instead. Therefore, following current practice, we will assume we
can remove the conditional means of our returns. Let
$\mu_t(\theta)=E[y_t | \Ic_{t-1}]=: E_{t-1}[y_t]$ be the
conditional mean vector of $y_t$. It depends on a vector of
parameters $\theta\in\Theta$. We define a ``detrended'' series
$(z_t)_{t\in\ZZ}$ by $$ y_t = \mu_t(\theta) +z_t,\;\;
E_{t-1}[z_t]=0.$$ For convenience, the conditional mean
$\mu_t(\theta)$ is assumed to be measurable with respect to
$\sigma(z_{t-1},z_{t-2},\ldots)$. Therefore,
$\Ic_t=\sigma(y_t,y_{t-1},\ldots) = \sigma(z_t,z_{t-1},\ldots)$.
Let us denote by $H_t$ the variance-covariance
matrix of the $t$-observations, conditionally on $\Ic_{t-1}$:
$Var(y_t|\Ic_{t-1})=Var(z_t|\Ic_{t-1}):=H_t.$
As usual with DCC-type models, we split the variance-covariance
matrix $H_t$ between volatility terms on one side (in $D_t$), and
correlation coefficients on the other side (in $R_t$):
\begin{equation}
H_t=D_t^{1/2} R_t D_t^{1/2}, \;\;D_t=diag(h_{1,t},...,h_{m,t}),
\label{Htdef}
\end{equation}
where $h_{k,t}$ denotes the ``instantaneous variance'' of
the return $y_{k,t}$ (or $z_{k,t}$, equivalently), conditionally on
$\Ic_{t-1}$.
We assume GARCH-type models on every
margin, but with potential cross-effects between all these volatilities:
\begin{equation}
Vecd(D_t) = V_0 + \sum_{i=1}^r
A_i .Vecd(D_{t-i}) + \sum_{j=1}^s B_j. \vec{z}_{t-j},
\label{Dtdef}
\end{equation}
for some deterministic non-negative matrices $(A_i)_{i=1,\ldots,r}$ and
$(B_j)_{j=1,\ldots,s}$, and for a positive vector $V_0$ in $\RR^m$. We will set
$A_i:=[a^{(i)}_{k,l}]_{1\leq k,l \leq m}$, $i=1,\ldots,r$, and
$B_j:=[b^{(j)}_{k,l}]_{1\leq k,l \leq m}$, $j=1,\ldots,s$.
\mds
Let us introduce the vector of so-called ``standardized residuals''
$\varepsilon_t := D_t^{-1/2}z_t$. Obviously, $E_{t-1}[\varepsilon_t]=0$ and $E_{t-1}[\varepsilon_t \varepsilon_t']=R_t$.
We impose that $M^{1/2}$ is positive definite for any positive definite matrix $M$. In this case, the
square root of $R_t$ is uniquely defined: see Serre (2010), Theorem 6.1. This will be our convention throughout the article.
\mds
The dynamics of correlations are given by the traditional Dynamic Conditional Correlation specification:
\begin{equation}
R_t=diag(Q_t)^{-\frac{1}{2}} Q_t diag(Q_t)^{-\frac{1}{2}}, \label{RtDCC}
\end{equation}
where the sequence of matrices $(Q_t)_{t\in\ZZ}$ satisfies
\begin{equation}
Q_t= W_0 + \sum_{k=1}^\nu M_k Q_{t-k} M_k' +
\sum_{l=1}^\mu N_l \eps_{t-l} \eps_{t-l}' N_l',
\label{Qtdef}
\end{equation}
for some deterministic matrices $(M_k)_{k=1,\ldots,\nu}$ and
$(N_l)_{l=1,\ldots,\mu}$, and for a positive definite constant
matrix $W_0$. Obviously, when such a sequence $(Q_t)_{t\geq -\nu}$ is initialized with $\nu$ non negative definite (possibly null) matrices, every $Q_t$, $ t\geq 0$, will be definite positive. In Theorem~\ref{thMoments}, we prove that a ``doubly infinite'' stationary sequence $(Q_t)_{t\in \ZZ}$ of definite positive matrices exists and which satisfies~(\ref{Qtdef}).
\mds
We will set $M_k:=[m^{(k)}_{p,q}]_{1\leq p,q \leq m}$, $k=1,\ldots,\nu$, and $N_l:=[n^{(l)}_{p,q}]_{1\leq p,q \leq m}$, $l=1,\ldots,\mu$. In practice, the positive matrix $W_0$ (or the constant vector $Vech(W_0)$ in $\RR^{m^*}$ equivalently) is a parameter that has to be estimated, most often during the first stage.
\mds
Aielli (2013) noticed that the estimation of the unknown matrix $W_0$ is not straightforward, because it cannot be deduced trivially from the unconditional correlation between the standardized residuals $\eps_t$. Therefore, he introduced a new variety of DCC-GARCH models (called cDCC), where~(\ref{Qtdef}) is replaced by
\begin{equation}
Q_t= W_0 + \sum_{k=1}^\nu M_k Q_{t-k} M_k' +
\sum_{l=1}^\mu N_l diag(Q_{t-l})^{-1/2} \eps_{t-l} \eps_{t-l}' diag(Q_{t-l})^{-1/2} N_l'.
\label{QtdefAielli}
\end{equation}
Under this new assumption, cDCC can be seen as a particular BEKK model (Engle and Kroner, 1995). Therefore, Aielli obtained the
existence of strictly and/or weakly stationary solutions, applying the conditions of Boussama, Fuchs, and Stelzer (2011) on BEKK processes. Actually, Aielli's model~(\ref{QtdefAielli}) is a smart but not intuitive ``ad-hoc'' specification. Its main justification appears as essentially technical, to avoid the non-linear feature of Engle's original DCC model~(\ref{Qtdef}). Under the standard latter specification, DCC models can no longer be rewritten as BEKK models and other techniques have to be found. In this paper, we obtain similar results to Aielli (2013), but by keeping the original specification of DCC models and without relying on another surrounding family of processes.
\subsection{DCC as Markov chains}
\label{DCCMarkov}
Actually, it is possible to rewrite the previous DCC model as a Markov chain, that looks like an AR(1) process. This rewrite will become a crucial tool when studying stationary solutions hereafter. Set
\begin{equation}
X_t := (X_t^{(1)},X_t^{(2)},X_t^{(3)},X_t^{(4)})',
\label{defXt}
\end{equation}
where
$$ X_t^{(1)}:= (Vecd(D_t),\ldots, Vecd(D_{t-r+1}))',$$
$$ X_t^{(2)} := (\vec{z}_t,\ldots,\vec{z}_{t-s+1})',$$
$$ X_t^{(3)}:= (Vech(Q_t),\ldots, Vech(Q_{t-\nu+1}))',\;\text{and}$$
$$ X_t^{(4)} := (Vech(\eps_t\eps_t'),\ldots,Vech(\eps_{t-\mu+1}\eps_{t-\mu+1}'))'.$$
The dimensions of the four previous random vectors are $rm$, $sm$,
$\nu m^*$ and $\mu m^*$ respectively. Their sum, the dimension of
$X_t$, is denoted by $d$. With simple block matrix calculations, random matrices $(T_t)$ and a vector process
$(\zeta_t)$ exist, such that the dynamics of $X_t$, any solution of the DCC model, may be rewritten as
\begin{equation}
X_t = T_t. X_{t-1}+\zeta_t,
\label{defMarkov}
\end{equation}
for any $t$. We will write the block matrix $T_t:=[T_{ij,t}]_{1\leq i,j\leq 4}$ with convenient
random matrices $T_{ij,t}$.
\mds
Knowing~(\ref{defMarkov}), the underlying process $(X_t)$ can be
seen as a vectorial autoregressive of order one, but with random
matrix-coefficients $(T_t)$.
Let us detail the AR(1) form of~(\ref{defMarkov}):
\begin{itemize}
\item set $T_{1k,t}=0$ when $k=3,4$,
$$ T_{11,t}:=
\left[
\begin{matrix}
A_1 & A_2 & \cdots & \cdots & A_r \\
I_m & 0_m & \cdots & \cdots & 0_m \\
0_m & I_m & 0_m & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0_m & \cdots & 0_m & I_m & 0_m
\end{matrix}
\right],\; \text{and}\;\,
T_{12,t}:= \left[
\begin{matrix}
B_1 & B_2 & \cdots & \cdots & B_s \\
0_m & \cdots & \cdots & \cdots & 0_m \\
\vdots & & & & \vdots \\
\vdots & & & & \vdots \\
0_m & \cdots & \cdots & \cdots & 0_m
\end{matrix}
\right]. $$
\item
We deduce from Equation~(\ref{Dtdef}) that
\begin{equation}
D_t \vec{\eps}_t = \vec{\eps}_t \odot Vecd(D_t) =\vec{z}_t = \vec{\eps}_t \odot V_0 + \sum_{i=1}^r
\vec{\eps}_t \odot A_i .Vecd(D_{t-i}) + \sum_{j=1}^s \vec{\eps}_t \odot B_j .\vec{z}_{t-j}.
\label{Dtdef2}
\end{equation}
Let us set $T_{23,t}=T_{24,t}=0$,
$$ T_{21,t}:=
\left[
\begin{matrix}
\vec{\eps}_t \odot A_1 & \vec{\eps}_t \odot A_2 & \cdots & \cdots & \vec{\eps}_t \odot A_r \\
0_m & \cdots & \cdots & \cdots & 0_m \\
\vdots & & & & \vdots \\
0_m & \cdots & \cdots & \cdots & 0_m
\end{matrix}
\right],\; \text{and} $$
$$
T_{22,t}:= \left[
\begin{matrix}
\vec{\eps}_t \odot B_1 & \vec{\eps}_t \odot B_2 & \cdots & \cdots & \vec{\eps}_t \odot B_s \\
I_m & 0_m & \cdots & \cdots & 0_m \\
0_m & I_m & 0_m & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0_m & \cdots & 0_m & I_m & 0_m
\end{matrix}
\right] $$
\item
Clearly, matrices $\tilde M_k$, $k=1,\ldots,\nu$, exist such that
$$ Vech(M_k Q_{t-k} M_k')=\tilde{M}_k .Vech(Q_{t-k}).$$
Similarly, matrices $\tilde N_l$, $l=1,\ldots,\mu$, exist such that
$$ Vech(N_l \eps_{t-l} \eps_{t-l}' N_l')=\tilde{N}_l .Vech(\eps_{t-l}.\eps_{t-l}').$$
It is possible to explicitly write the previous matrices $\tilde M_k$ and $\tilde N_l$.s
Indeed, with the notations of Subsection~\ref{notations}, $ \tilde M_k = [\tilde
m_{u,v}^{(k)}]_{1\leq u,v \leq m^*}$ where
$$ \tilde m_{u,v}^{(k)} = m^{(k)}_{\phi_1(u),\phi_1(v)}m^{(k)}_{\phi_2(u),\phi_2(v)}.$$
Similarly,
$ \tilde N_l = [\tilde
n_{u,v}^{(l)}]_{1\leq u,v \leq m^*}$ and
$ \tilde n_{u,v}^{(l)} = n^{(l)}_{\phi_1(u),\phi_1(v)}m^{(l)}_{\phi_2(u),\phi_2(v)} .$
Then, set $T_{31,t}=T_{32,t}=0$,
$$ T_{33,t}:=
\left[
\begin{matrix}
\tilde{M}_1 & \tilde{M}_2 & \cdots & \cdots & \tilde{M}_\nu \\
I_{m^*} & 0_{m^*} & \cdots & \cdots & 0_{m^*} \\
0_{m^*} & I_{m^*} & 0_{m^*} & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0_{m^*} & \cdots & 0_{m^*} & I_{m^*} & 0_{m^*}
\end{matrix}
\right] ,\; \text{and}\;
T_{34,t}:=
\left[
\begin{matrix}
\tilde{N}_1 & \tilde{N}_2 & \cdots & \cdots & \tilde{N}_\mu \\
0_{m^*} & \cdots & \cdots & \cdots & 0_{m^*} \\
\vdots & & & & \vdots \\
0_{m^*} & \cdots & \cdots & \cdots & 0_{m^*}
\end{matrix}
\right]. $$
\item
$T_{4k,t}=0$, $k=1,2,3$, and define the $\mu m^* \times \mu m^*$ matrix
$$T_{44,t}:=
\left[
\begin{matrix}
0_{m^*} & 0_{m^*} & \cdots & \cdots & 0_{m^*} \\
I_{m^*} & 0_{m^*} & \cdots & \cdots & 0_{m^*} \\
0_{m^*} & I_{m^*} & 0_{m^*} & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0_{m^*} & \cdots & 0_{m^*} & I_{m^*} & 0_{m^*}
\end{matrix}
\right] .$$
\end{itemize}
Moreover, rewrite
$\zeta_t=(\zeta_t^{(1)},\zeta_t^{(2)},\zeta_t^{(3)},\zeta_t^{(4)}),$
where, with obvious sizes, these vectors are
$$ \zeta_t^{(1)}=(V_0,0_m,\ldots,0_m)',\;\; \zeta_t^{(2)}=(\vec{\eps}_t \odot V_0 ,0_m,\ldots,0_m)',$$
$$ \zeta_t^{(3)}=(Vech(W_0),0_{m^*},\ldots,0_{m^*})',\;\text{and}\;\; \zeta_t^{(4)}=(Vech(\eps_t\eps_t'),0_{m^*},\ldots,0_{m^*})'.$$
\mds
Since $D_t$, $Q_t$ and $R_t$ are
$\Ic_{t-1}-$measurable, it is easy to see that the filtration induced by the observations is the natural filtration of $(X_t)$:
$\sigma(X_t,X_{t-1},\ldots)=\Ic_t$ for all $t$. From now on, we consider $(\Ic_t)$ as the filtration that is generated by $(X_t)$. This means that $E_{t-1}[Z]=E[Z | \Ic_{t-1}]=E[Z | X_{t-1},X_{t-2},\ldots]$, for any random vector $Z$. And a process $(Z_t)$ is said to be (one-order, implicitly) $\Ic$-Markov if the law of $Z_t$ given $\Ic_{t-1}$ is the
law of $Z_t$ given $X_{t-1}$.
\mds
Intuitively, the sequence $(X_t)$ is $\Ic$-Markov because it is the
case for the processes $(\zeta_t)$ and $(T_t)$ themselves. To prove this formally, we need an assumption concerning the data generating process (DGP) of $(z_t)$.
\mds
Let us define the $t$-vector of innovations by
\begin{equation}
\eta_t := R_t^{-1/2} \varepsilon_t = R_t^{-1/2} D_t^{-1/2} z_t.
\label{defEta}
\end{equation}
Note that $E_{t-1}[\eta_t]=0$ and $E_{t-1}[\eta_t\eta_t']=I_m$ by construction.
The definition of
these innovations implies that, for every $t$,
$\sigma(\eta_j, j\leq t) \subset \sigma(\eps_j, j\leq t) \subset \Ic_t .$
Nonetheless, we will not establish whether there are equalities between
the latter filtrations. Technically speaking, this would be
equivalent to stating the invertibility of the underlying process.
\mds
{\it\bf Assumption A0:}
$(\eta_t)_{t\in \ZZ}$ possesses the Markov property with respect to the filtration $\Ic$. In particular, $E[\eta_t | \Ic_{t-1}]=
E[\eta_t |X_{t-1}]$ for every $t$.
\mds
Obviously, the latter assumption is satisfied if $(\eta_t)_{t\in \ZZ}$ is a sequence of identically distributed and mutually
independent random vectors, with $E[\eta_t]=0$ and $E[\eta_t \eta_t']=I_m$.
For instance, if the random vectors $\eta_t$ are standardized Gaussian and mutually independent, the process $(z_t)_{t\in\ZZ}$ is conditionally Gaussian, a standard case in practice.
\mds
\begin{proposition}
\label{prop_Markov}
Under A0, the process $(X_t)$ is Markov of order one with respect to its natural filtration.
\end{proposition}
See the proof in the appendix.
\begin{remark}
It would be possible to define a slightly different DGP for the DCC model above: consider a $\Ic$-martingale difference (i.i.d., for instance) sequence $(e_t)_{t\in\ZZ}$ and set
$z_t = H_t^{1/2} e_t$, $E_{t-1}[e_t]=0$, $E_{t-1}[e_t e'_t]=I_m$. In the latter case, the process $(\eta_t)$ above would be defined by $\eta_t = R_t^{-1/2} D_t^{-1/2} H_t^{1/2} e_t$. Then, it is easy to check that $(\eta_t)$ is $\Ic$-Markov and
is a martingale difference. In other words, A0 would apply in such circumstances.
\end{remark}
\section{Stationarity of DCC models}
\subsection{Existence of stationary DCC solutions}
\label{MainRes}
The AR dynamics of $X_t$ were defined above thanks to $T_t$ and $\zeta_t$, which will be stochastic only through $\eps_t$, i.e. through the $t$-innovation $\eta_t$ and the $\Ic_{t-1}$-measurable matrix $R_t$. This creates a major difficulty in proving the existence of stationary solutions. In particular, this means that $T_t$ depends on some components of $X_t$. Therefore, it will be difficult to find explicit expressions
like $X_t = f(\eta_t,\eta_{t-1},\ldots)$ for some deterministic and measurable function $f$, because the link between $T_t$ and the past
innovations (or observations) is highly non-linear.
\mds
To obtain the existence of stationary solutions in the previous DCC model, we will invoke Tweedie's (1988) criterion. The latter result will provide the existence of an invariant probability measure for the Markov chain defined by~(\ref{defMarkov}). sThis technique has already been used in several papers in econometrics, notably Ling and McAleer (2003) or Ling (1999).
\mds
To get the stationarity conditions of $(z_t)$, we have to control the magnitude of the random matrix $T_t$, which depends on the random variables $\eps_{kt}^2$, $k=1,\ldots,m$. The mean of the latter variables is one, but they are not independent. This is in contrast with Ling and McAleer (2003). Moreover, unfortunately, the joint law of $\vec{\eps}_t$ is a function of $R_t$, i.e. a function of $X_{t-1}$. That is why we need
the following condition.s
\mds
{\it\bf Assumption E1:} For some $p\geq 1$, $E[\| \eta_t
\|^{2p}]<\infty$ and $ \rho\left(T^*\right)<1$, where
$$ T^*:=\sup_{\xx\in \RR^d} E[|T_t^{\otimes p}| \; | \; X_{t-1}=\xx ].$$
\mds
We reiterate that $T_t$ depends on $\vec{\eps}_t$, that $\eps_t=R_t^{1/2} \eta_t$, and that the components of $\eta_t$ are uncorrelated. As such, the coefficients of $ T^*$ are finite because all of the coefficients of $R_t$ are less than one (in absolute values). When there are no correlation dynamics, the matrices $M_k$ and $N_l$ are zero and we recover CCC models. In the latter case, our Assumption E1 is reduced to the main assumption of Ling and McAleer (Theorem 2.2) that was stated for vectorial ARMA-GARCH models.
\mds
{\it\bf Assumption E2:}
The law of $\eta_t$ given that $X_{t-1}=\xx$ is absolutely continuous with respect to the Lebesgue measure, and its density is denoted by $f_{\eta_t}(\cdot |\xx)$, for every $\xx \in \RR^d$ and $t$.
The function $\xx \mapsto f_{\eta_t}(\eta | \xx)$ is continuous for every $\eta\in \RR^m$ and $t$.
There exists an integrable function $H$ s.t. $ \sup_t\sup_{\xx\in \RR^d} f_{\eta_t}(\eta |\xx) \leq H(\eta)$
for every $\eta\in \RR^m$.
Moreover, $\sup_t E[\| \eta_t \|^{2p} \, | X_{t-1}=\xx] \leq \bar h(\|\xx \|)$, for some function $\bar h$ that satisfies
$\lim_{v\rightarrow +\infty} \bar h(v)/v^\gamma =0$ for every $\gamma >0$.
\mds
The latter technical assumption is trivially satisfied when $(\eta_t)$ is an i.i.d. sequence of random vectors s.t. $E[\| \eta_t \|^{2p}]<+\infty$.
Otherwise, E2 provides some constraints insofar as the law of $\eta_t$ depends on the past values of the DCC process.
Similar conditions may appear in the literature about the non-parametric estimation of conditional expectations.
However, most of them relate to the boundedness of $\bar h$ and/or its derivatives (as in Assumption 3 in Newey 1997, for instance), or to the moments of $\bar h$ (as Assumption 1 in Donald et al. 2003, for instance). Clearly, E2 is weaker than such assumptions.
\mds
\begin{theorem}
\label{thMoments} Under the assumptions A0 and E1-E2, the process $(z_t,D_t,R_t)$
as defined by Equations~(\ref{Htdef}),~(\ref{Dtdef}),~(\ref{RtDCC})
and~(\ref{Qtdef}), possesses a strictly stationary solution. The
latter process is measurable with respect to the $\sigma$-field $\Ic$ induced
by the observations. Moreover, the $2p$-th moments of a solution $(z_t)$ are finite.
\end{theorem}
{\bf Example 1:} In practice and for the sake of parsimony, it is usual to assume diagonal-type DCC models, where all the
parameter matrices are diagonal, assuming no ``cross-effects''
in terms of volatilities and/or correlations. This means the
non-negative real numbers $a^{(i)}_u$, $b^{(j)}_u$, $m^{(k)}_u$ and $n^{(l)}_u$,
$u=1,\ldots,m$, exist such that
$$ A_i = diag(a^{(i)}_1,\ldots,a^{(i)}_m),\, i=1,\ldots,r,\; B_j=diag(b^{(j)}_1,\ldots,b^{(j)}_m),\, j=1,\ldots,s,$$
$$ M_k = diag(m^{(k)}_1,\ldots,m^{(k)}_{m^*}),\, k=1,\ldots,\nu, \; N_l=diag(n^{(l)}_1,\ldots,n^{(l)}_{m^*}),\, l=1,\ldots,\mu.$$
The associated matrices $\tilde M_k$ and $\tilde N_l$ are also diagonal.
Set $\tilde M_k = diag(\tilde m^{(k)}_l)_{1\leq l \leq m^*}$, and check that $\tilde m^{(k)}_l=m_{\phi_1(l)}^{(k)} m_{\phi_2(l)}^{(k)}$.
Now, let us specify the previous Assumption E1 when $p=1$.
\mds
Since $E[\vec{\eps}_{kt}\; | \; X_{t-1}=\xx ]=1$ for every index $k$,
$T^*$ is simply $|T_t|$, replacing $\vec{\eps}_t$ by one. Denote
by $P^*$ the characteristic polynomial of $T^*$, i.e.
$P^*(\lambda)=\text{Det}(T^* - \lambda Id)$. It can be seen easily
that two polynomials $P_1^*$ and $P_2^*$ s.t.
$ P^*(\lambda)= P_1^*(\lambda) P_2^*(\lambda)$ exist. Here, $P_1^*$ denotes the
characteristic polynomial of the block-matrix $[|T_{ij,t}|]_{1,\leq
i,j\leq 2}$, replacing $\vec{\eps}_t$ by one. $P_2^*$ is the
characteristic polynomial of the previous matrix $|T_{33,t}|$. Tedious, but relatively uncomplicated,
algebraic calculations provide
$$ P_1^*(\lambda) = \pm \lambda^{\pi_1} \Prod_{k=1}^m \left(\sum_{i=1}^r a_k^{(i)}\lambda^{r+s-i} + \sum_{j=1}^s
b_k^{(j)} \lambda^{r+s-j} -\lambda^{r+s}\right),$$
$$ P_2^*(\lambda) = \pm \lambda^{\pi_2} \Prod_{l=1}^{m^*} \left(\sum_{k=1}^\nu \tilde m^{(k)}_l \lambda^{\nu-k} -\lambda^{\nu}\right),$$
for some integers $\pi_1$ and $\pi_2$.
Let $\lambda_0$ be a non-zero root of $P^*$. If $\lambda_0$ is a root of $P_1^*$ then there exists an index $k\in\{1,\ldots,m\}$ such that
$ \sum_{i=1}^r a_k^{(i)}\lambda_0^{r+s-i} + \sum_{j=1}^s
b_k^{(j)} \lambda_0^{r+s-j} =\lambda_0^{r+s} .$
If $|\lambda_0|\geq 1$, this implies
$ 1 \leq \sum_{i=1}^r a^{(i)}_k + \sum_{j=1}^s b^{(j)}_k.$
On the other side and similarly, if $\lambda_0$ is a root of $P_2^*$ and if $|\lambda_0|\geq 1$, then there exists $l\in\{1,\ldots,m^*\}$ s.t.
$ 1 \leq \sum_{k=1}^\nu |\tilde m^{(k)}_l| .$
In other words, a sufficient condition to fulfill Assumption E1 is
\begin{equation}
\sup_{k=1,\ldots,m} \sum_{i=1}^r a^{(i)}_k + \sum_{j=1}^s b^{(j)}_k <1,\;\; \text{and}\;\;
\sup_{l=1,\ldots,m^*}\sum_{k=1}^\nu |\tilde m^{(k)}_l| <1 .
\label{Th1DiagDCCCond}
\end{equation}
\mds
Nonetheless, to apply Theorem~\ref{thMoments} in the general case, it may be hard to check the condition
on the spectral radius of $T^*$. This is due to
the analytical complexity of $T_t^{\otimes p}$, $p>1$, or to the calculation of its eigenvalues, even when $p=1$. In the next theorem,
we provide more explicit conditions in the case $p=1$, i.e. so that the second-order moments of $(z_t)$ are finite. These conditions ensure that E1 will be satisfied.
In other words, the conditions will be stronger than E1, but they may be more practical.
Indeed, it is often important to obtain sufficient conditions that can be written explicitly in terms of the model parameters, for instance for inference purposes (e.g. the optimization stage to get QML estimates).
\mds
Let us consider $\Nc$ (resp. $\Nc^*$) an arbitrary norm for vectors in $\RR^m$ (resp. $\RR^{m^*}$). Denote by $\|\cdot \|_{\Nc}$
and $\|\cdot \|_{\Nc^*}$ the associated norms for matrices.
\begin{theorem} If
\label{thSimple}
\begin{equation}
\sum_{i=1}^r \| A_i \|_{\Nc} + \sum_{j=1}^s \| B_j \|_{\Nc} <1,\;\; \text{and}
\label{cond1ThSimple}
\end{equation}
\begin{equation}
\sum_{k=1}^\nu \| \tilde M_k \|_{\Nc^*} <1,
\label{cond2ThSimple}
\end{equation}
then Assumption E1 is satisfied with $p=1$.
\end{theorem}
Note that the conditions of Theorem~\ref{thSimple} do not depend on
the matrices $N_l$, $l=1,\ldots,\mu$, which is a relatively unexpected result.
Once they are satisfied, and under A0 and E2, Theorem~\ref{thMoments} applies.
\mds
By choosing $\Nc$ as the maximum norm for vectors, it can easily be checked that $\|A_i\|_{\Nc}=\sup_{p=1,\ldots,m} \sum_{q=1}^m a^{(i)}_{p,q}$, and similarly with the matrices $B_j$.
Alternatively, we can choose $\Nc(\xx)=\|x\|_2$, that induces the spectral norm $\|A_i \|_{\Nc}=\|A_i \|_s$. Obviously,
we can choose these norms for $\Nc^*$ and the matrices $\tilde M_k$.
\mds
It is often of value to assume that the Markov chain is initialized at $t=0$ by drawing $X_0$ following its stationary law. Introducing the filtration $\Ic_t^*:=\sigma(X_0,z_1,\ldots,z_t)$, we can easily see that the DCC solution is now measurable with respect to the
$\sigma$-field induced by the innovations and the initial value, because, whenever $t>0$,
$$ \sigma(X_0,z_1,\ldots,z_t)=\sigma(X_0,\eps_1,\ldots,\eps_t)=\sigma(X_0,\eta_1,\ldots,\eta_t).$$
\mds
{\bf Example 1 (Continued):} Consider a diagonal-type DCC model and maximum norms for vectors.
In this case, the condition~(\ref{cond1ThSimple}) becomes
$$ \sum_{i=1}^r \sup_{l=1,\ldots,m} a_{l}^{(i)} + \sum_{j=1}^s \sup_{l=1,\ldots,m} b_{l}^{(j)} <1,$$
and the condition~(\ref{cond2ThSimple}) is
$\sum_{k=1}^\nu \sup_{l=1,\ldots,m^*}|\tilde m^{(k)}_l| <1.$
These two conditions are stronger than~(\ref{Th1DiagDCCCond}), as expected.
\mds
{\bf Example 2:}
To reduce the number of free parameters even further, scalar-DCC models are often introduced. In
this case, all the unknown matrices are simply products of a
scalar and an identity matrix:
$$ A_i = a^{(i)}I_m,\, i=1,\ldots,r,\; B_j=b^{(j)}I_m,\, j=1,\ldots,s,$$
$$ M_k = m^{(k)}I_m,\, k=1,\ldots,\nu, \; N_l=n^{(l)}I_m,\, l=1,\ldots,\mu.$$
Such models are very popular, because they allow the number of free parameters to be drastically reduced. With obvious notations,
the conditions of Theorem~\ref{thMoments} and~\ref{thSimple} are the same as above:
\begin{equation}
\sum_{i=1}^r a^{(i)} + \sum_{j=1}^s b^{(j)} <1,\; \text{and}\; \sum_{k=1}^\nu |m^{(k)} |^2 <1.
\label{scalarDCCcond}
\end{equation}
In passing, we recover the usual (second-order and strict)
conditions of stationarity for GARCH-type models:
$$0\leq a^{(i)},b^{(j)}\leq 1,\;\text{and}\;\;\sum_{i=1}^r a^{(i)} + \sum_{j=1}^s b^{(j)} <1.$$
\mds
\subsection{Uniqueness of stationary DCC solutions}
\label{Unicity}
Even if stationary solutions of the DCC model do exist, we are not initially sure {\it a priori} that they are unique. Besides its theoretical interest, this problem has practical implications. For instance, for any process, the convergence of simulated trajectories towards the same stationary law, independently of the initialization stage, is a desirable feature. Moreover, the uniqueness of invariant measures of a Markov process implies the ergodicity of the stationary solution (see Douc et al. 2014, Corollary 7.17). This is particularly important for inference purposes. Indeed, the estimation of DCC models is typically based on M-estimates (Quasi Maximum Likelihood, for instance). These techniques rely heavily on uniform Laws of Large Numbers, that are most often deduced from the ergodicity of the process. The conditions for identifiability and consistency rely on some expectations with respect to the underlying invariant measure of the given stationary process. If, for a given set of parameters, several invariant measures exist, then it becomes difficult to check such conditions. Finally, with several underlying invariant measures, we cannot exclude the possibility of switches from one stationary trajectory to another, disturbing the econometric analysis (stationarity tests, statistical uncertainty around estimates, etc).
\mds
Unfortunately, this uniqueness is not given ``for free'' by Tweedie's Lemma~\ref{TweedieLemma}. Moreover, the usual arguments concerning
the uniqueness of stationary GARCH-type solutions do not apply here. Indeed, under the Markov-chain specification given by Equation~(\ref{defMarkov}), the matrix $T_t$ is itself a function of the random vector $X_t$ through the $\vec{\eps}_t$ factors. This is a
major difference with the CCC case, and we need to find another strategy. In this section, we provide some uniqueness results under some more or less restrictive assumptions.
\mds
Now, we will consider only stationary solutions of the DCC model, as given in Section~\ref{MainRes}.
We know that such solutions exist under the (sufficient) conditions of Theorem~\ref{thMoments} or~\ref{thSimple}, but it is not necessary to impose
such conditions from the outset. Obviously, we will need other technical assumptions.
\mds
{\it\bf Assumption U0:}
The sequence of innovations $(\eta_t)_{t\in \ZZ}$ is highly stationary and ergodic.
\mds
{\it\bf Assumption U1:} $\| T_{33} \|_s<1$.
\mds
The matrix $T_{33}$ has been introduced in
Subsection~\ref{DCCMarkov}, under the name $T_{33,t}$. Since
$T_{33,t}$ does not depend on time, we have removed the index $t$ here.
\mds
Thanks to the latter assumptions, we will be able to bound
$\|Q_t\|_{\max}$ from above by a stationary process $(q_t)$, and from below by a constant. Moreover, $\lambda_1(Q_t)$ will be bounded from
below. These tools will be crucial in proving the uniqueness of
stationary DCC solutions. To do so, let us introduce some intermediate quantities.
\begin{itemize}
\item The process $(q_t)$, defined by
$$q_t :=\frac{\| Vech(W_0) \|_2}{1- \| T_{33}\|_s} + \sqrt{\frac{m^3(m+1)}{2}} \sum_{l=1}^\mu \| \tilde N_l\|_{s}\xi_{t-l},$$
where
$\xi_t := \sum_{k=0}^{+\infty} \| T_{33}\|_s^k \|\eta_{t-k}\|_2^2.$
\item The constants $C_\lambda :=\lambda_1(W_0)$ and $C_q:=\min_{i=1,\ldots,m}(W_0)_{ii}$.
\item The constants
$$C^*_\lambda := \frac{ \lambda_1(W_0)}{1- \sum_{k=1}^\nu (m^{(k)})^2},\; \text{and} \;
C_q^*:=\frac{\min_{i=1,\ldots,m}(W_0)_{ii}}{1- \sum_{k=1}^\nu (m^{(k)})^2}\cdot$$
\item
$\kappa = \max(\nu,\mu)$ and, for every $j=1,\ldots,\kappa$, set
$$ \beta_{j,t}:= \1 ( j\leq \nu) \| M_j \|_s^2 + \1(j\leq \mu) \| N_j \|_s^2
\frac{4(2m+1)m^{1/2} }{\sqrt{C_\lambda}C_q} \|\eta_t \|_2^2
\sqrt{q_t}.$$
\end{itemize}
\mds
Let $N_t^*$ be the $(\kappa,\kappa)$-squared random matrix
$$ N_t^* :=
\left[
\begin{matrix}
\beta_{1,t} & \beta_{2,t} & \cdots & \cdots & \beta_{\kappa,t} \\
1 & 0 & \cdots & \cdots & 0 \\
0 & 1 & 0 & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0 & \cdots & 0 & 1 & 0
\end{matrix}
\right].$$
Note that the sequences $(\xi_t)$, $(q_t)$ and $(N_t^*)$ are stationary and ergodic because any $\xi_t$, $q_t$ or $N_t^*$ is a measurable function of the innovations $(\eta_t)$ that are stationary and ergodic under Assumption U0.
\mds
{\it\bf Assumption U2:} $E[\ln^+ \| N_t^*\|]<\infty$ and the top Lyapunov exponent of the sequence $(N_t^*)$, defined by
$\gamma_N:=\lim_{t\rightarrow +\infty} t^{-1}E[\ln (\| N_1^*N_2^*\ldots N_t^*\|) ]$, is strictly negative.
\mds
Such conditions are standard in the GARCH
literature (see Francq and Zako\"{\i}an, 2010, Section 2.2.2. for instance). Note that $\gamma_N \leq E[\ln \| N_1^*\|]$ for any norm $\|\cdot \|$.
\mds
Actually, the technical assumptions U1-U2 above will ensure the uniqueness of $(\eps_t)$, $(Q_t)$ and $(R_t)$ only. To get the uniqueness of $(D_t)$ and then of $(z_t)$ itself, we need a last assumption: with the notations of
Subsection~\ref{DCCMarkov}, set
$$ \bar{T}_t := \left[
\begin{array}{cc}
T_{11,t} & T_{12,t} \\
T_{21,t} & T_{22,t}
\end{array}
\right],\;
\text{and}\; \bar{T}^* = E[\bar T_t].$$
Note that $\bar{T}^*$ does not depend on any particular sequence $(\eps_t)$ nor $t$, because $E[\eps^2_{kt}]=1$ for every $k$.
\mds
{\it\bf Assumption U3:} $\rho(\bar T^*)<1$
\mds
\begin{theorem}
\label{ThUnicity} Under A0 and U0-U3, a strictly stationary
solution of the DCC model is unique and ergodic, given a sequence $(\eta_t)$.
\end{theorem}
\mds
The latter result can be strengthened in the following particular case, which is commonly encountered in the literature.
{\it\bf Assumption U4:} The underlying DCC model is ``partially''
scalar, i.e. scalars $m^{(k)}$ exist such that $M_k = m^{(k)}
I_m$ for all $k=1,\ldots,\nu$. Moreover, $\rho(M^*)<1$ by setting
$$ M^* :=
\left[
\begin{matrix}
(m^{(1)})^2 & (m^{(2)})^2 & \cdots & \cdots & (m^{(\nu)})^2 \\
1 & 0 & \cdots & \cdots & 0 \\
0 & 1 & 0 & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0 & \cdots & 0 & 1 & 0
\end{matrix}
\right].$$
\mds
Obviously, U4 is not mandatory to get our uniqueness result, even
if it allows the technical condition U2 to be weakened most often, by lowering the $\beta_{j,t}$ terms. In every
case, this ``partially'' scalar case encompasses the common
practice of scalar DCC (or scalar multivariate GARCH) models.
\begin{corollary}
\label{ThUnicityPL}
Under A0 and U0-U4, a strictly stationary
solution of the DCC model is unique and ergodic, given a sequence $(\eta_t)$, replacing $C_\lambda$ (resp. $C_q$) by
$C^*_\lambda$ (resp. $C_q^*$) in U2.
\end{corollary}
\mds
{\bf Example 2 (Continued):}
In the case of scalar DCC models of order one, it is easy to specify the conditions above. Here, $r=s=\nu=\mu=1$,
$$ A_1 = a^{(1)} I_m, \; B_1=b^{(1)} I_m, \; M_1 = m^{(1)}I_m, \; N_1=n^{(1)}I_m.$$
Assumptions U1 and U4 are equivalent and mean $|m^{(1)}|<1$.
Assumption U4 is fulfilled if $E[\ln \| N_1^*\|_{\max}]<0$, or if
\begin{equation}
\label{Ex2Unicity}
E\left[ \ln\left( (m^{(1)})^2 + (n^{(1)})^2
\frac{4(2m+1)m^{1/2} }{\sqrt{C_\lambda}C_q} \|\eta_t \|_2^2
\sqrt{q_t} \right) \right] <0.
\end{equation}
This expectation could be easily evaluated by simulation, by noting that $\eta_t$ and $q_t$ are independent.
Finally,
$$ \bar T^* = \left[
\begin{array}{cc}
a^{(1)} & b^{(1)} \\
a^{(1)} & b^{(1)}
\end{array}
\right] \otimes I_m
.$$
Through elementary algebra, it can checked that the characteristic function of $\bar T^*$ is the function $x\mapsto (-x)^m(a^{(1)}+ b^{(1)} - x)^m$. Then Assumption U3 means $a^{(1)}+ b^{(1)} <1$.
Therefore, as expected, the conditions required for stationary DCC solutions to be unique are more demanding than for them to simply exist, due to U2. Generally, the latter condition will be fulfilled more easily if $\sup_l \|N_l \|_s$ is ``a lot smaller'' than one, if $m$ is not ``too large'', and if the tails of $\eta_t$ are not ``too heavy''.
\mds
\section{Discussion and practical considerations}
\label{My_discussion}
Now, let us discuss the sufficient conditions to obtain the existence of stationary DCC solutions, as given in Theorems~\ref{thMoments} and~\ref{thSimple}.
The most explicit ones are~(\ref{cond1ThSimple}) and~(\ref{cond2ThSimple}). Since scalar DCC models are by far the most commonly used models in the literature, we focus on the conditions of the previous examples 1 and 2 above, particularly~(\ref{scalarDCCcond}). In the latter case, the conditions on the coefficients of the volatility process are those generally applied in the univariate GARCH literature. It can be proved they are necessary and sufficient for the existence of second-order and strictly stationary GARCH solutions (see Francq and Zako\"{\i}an, 2010, Theorem 2.6 and Remark 2.6). More interestingly, we can check empirically how tight the (now) new conditions on the coefficients of the $(Q_t)$ process are. In other words, in Example 2, is the constraint
$ \sum_{k=1}^\nu |m^{(k)} |^2 <1 $
close to a necessary and sufficient condition for generating stationary trajectories of $(z_t)$, $(R_t)$ and/or $(Q_t)$s?
\mds
For illustrative purposes, we have considered a very simple bivariate scalar DCC model of order one, given by
$$ h_{k,t}= v_{0} + a\, h_{k,t-1} + b \,\eps_{k,t}^2, \; k=1,2,$$
$$ v_0 = 1/4,\, a=0.8 ,\, b=0.1,$$
$$ Q_t = W_0 + (m^{(1)})^2 Q_{t-1} + (n^{(1)})^2 \eps_{t-1} \eps_{t-1}',$$
$$ W_0 = I_2/2 +ee'/2,\, e=[1,1]', $$
with our notations. The latter process is generated by i.i.d. innovations $(\eta_t)$ that are independent standard bivariate Gaussian vectors. We initialize
the process at $t=0$ with $Q_0=R_0=I_2$ and $h_{k,0}=1/2$, $k=1,2$.
\mds
In the paper, we have stated theoretically that the values of the coefficients of the matrices $N_l$, $l=1,\ldots, \mu$ (or the coefficients $n^{(l)}$ in the scalar case) do not matter to obtain the existence of stationary DCC solutions. We have verified this fairly counter-intuitive fact empirically: with the model above, different coefficients $n^{(1)}$
do not seem to modify the shape of the trajectories we generate, independently of the other parameters. Therefore, our experiments will lead with a fixed value $n^{(1)}=\sqrt{3}$.
Note that this value is larger than one and this could be seen ``naively'' as a source of non-stationarity.
\mds
As expected, the value one is key for $m^{(1)}$. When the latter is very close to one but less than one ($(m^{(1)})^2=0.999$ in our case),
we check that the simulated trajectories of $(z_t)$, $(Q_t)$ and $(R_t)$ look stationary, once the influences of the starting values have been forgotten (broadly speaking when $t\geq 4000$): see Figure~\ref{QtFig} (solid lines).
On the other side, when this auto-regressive parameter is larger than one, even by a small amount ($(m^{(1)})^2=1.001$ in our case), we observe that the $(Q_t)$ trajectories explode: see Figure~\ref{QtFig} (dashed lines). Apparently, this is not the case for the $(R_t)$ correlation coefficients. They tend towards some constant levels (Figure~\ref{RtFig}), but differ from one experiment to another one. This phenomenon is a consequence of the normalization stage~(\ref{RtDCC}), but it seems difficult to maintain that feature will happen for almost every trajectory and any DCC model. Indeed, by managing increasingly (very) high numbers with the $(Q_t)$ process, we cannot exclude the possibility of spurious unexpected $(R_t)$ behaviors. Moreover, we have checked that the $(z_t)$ trajectories that we generated in this case do not seem to exhibit non-stationary patterns visually. This is logical because our DCC model tends towards a CCC model in such situations.
Therefore, it is likely that modelers will be able to manage (i.e. evaluate and simulate numerically) DCC trajectories in practice,
even if~(\ref{cond2ThSimple}) or its generalizations are not satisfied. Actually, this task remains feasible as long as the numerical values of $(Q_t)$ are manageable (i.e. not too large) by our software. This is the case when the number of dates is not too large.
\mds
We have replaced Gaussian innovations $(\eta_t)$ by fat-tailed random vectors, to check to what extent this may be a source of instability.
The $\eta_{kt}$ components, $k=1,2$ and $t=0,\ldots,T$, have been drawn following mutually independent standardized Student laws with $\nu$ degrees of freedom, $\nu>0$.
We reiterate here that the $\eta_t$-moments of order $\nu$ or higher do not exist.
As long as $\nu \geq 2$ and $m^{(1)}<1$, we do not observe explosive patterns for $(z_t)$. The components of $R_t$ appear to become stationary, even if the
decrease in initial value effects is very slow when $\nu$ is close to two.
On the contrary, when $\nu< 2$, the processes $(Q_t)$ and $(R_t)$ are highly unstable. The former exhibit very spiky trajectories, while the latter often tend to be attracted
by the $1$ or $(-1)$ area.
Besides, some $(z_t)$ trajectories reach very high and unrealistic values ($10^{80}$ for instance).
In every case, when $m^{(1)}\geq 1$, the $(Q_t)$ trajectories explode and the $(R_t)$ ones tend to constant values that depend on each experiment. In such cases, the return process
$(z_t)$ may reach huge values, but only when $\nu \leq 2$, apparently. This analysis illustrates the necessity of considering innovations with finite second-order moments, and the fact that
higher-order moments are not mandatory to obtain stationary solutions.
\mds
Concerning the sufficient conditions that guarantee that stationary DCC solutions are unique, it is more difficult to evaluate their tightness
because they are more intricate and involve too many model characteristics. Nonetheless, with the simple scalar DCC model used in Example $2$, we observe that the
key condition~(\ref{Ex2Unicity}) will be more demanding when the number $m$ of underlyings increases. On the contrary, smaller values $|n^{(1)}|$ would help.
And the partially scalar case induces significantly less demanding conditions than the general case, because the values of the constants
$ C_\lambda^*$ and $C_q^*$ in the denominator are a lot higher than $C_\lambda$ and $C_q$ respectively. The effect of $|m^{(1)}|$ is ambiguous because it
appears in several quantities, especially $T_{33}$ and $C^*_\lambda$.
\mds
\section*{Acknowledgments}
\mds
{\rm We thank C. Francq and J.-M. Zako\"{\i}an for their valuable
remarks and discussions. Moreover, we are grateful to Eric Renault and two anonymous referees, who have proposed a number of ways of improving the article.
Finally, the authors thank the Labex ``Ecodec'' for its support.}
|
\partial{\partial}
\def \ \rm{ Id } { \ \rm{ Id } }
\def\mathcal M{\mathcal M}
\def\mathcal L{\mathcal L}
\def\mathcal R{\mathcal R}
\def\mathcal S{\mathcal S}
\def\Sc{{\rm R}}
\def\nabla{\nabla}
\def\gradg{\left| {{}^{{}^h}\!\nabla} g\right|}
\def\hGamma{ {{}^h \Gamma} }
\def\displaybreak[1]{\displaybreak[1]}
\newcommand{\tfrac{1}{2}}{\tfrac{1}{2}}
\DeclareMathOperator{\vol}{vol}
\def\,d\!\vol_h{\,d\!\vol_h}
\DeclareMathOperator{\Riem}{Riem}
\DeclareMathOperator{\Ric}{Ric}
\DeclareMathOperator{\Ricci}{Ric}
\DeclareMathOperator{\diag}{diag}
\DeclareMathOperator{\dist}{dist}
\DeclareMathOperator{\id}{id}
\def\text{\it{loc}}{\text{\it{loc}}}
\def\counterword#1{%
\ifthenelse{\ref{#1}=1}{one}{}%
\ifthenelse{\ref{#1}=2}{two}{}%
\ifthenelse{\ref{#1}=3}{three}{}%
\ifthenelse{\ref{#1}=4}{four}{}%
\ifthenelse{\ref{#1}=5}{five}{}%
\ifthenelse{\ref{#1}=6}{six}{}%
}
\def\rule{5cm}{1ex}{\rule{5cm}{1ex}}
\newtheorem{theorem}{Theorem}[section]
\newtheorem{lemma}[theorem]{Lemma}
\newtheorem{corollary}[theorem]{Corollary}
\newtheorem{conjecture}[theorem]{Conjecture}
\newtheorem{proposition}[theorem]{Proposition}
\theoremstyle{definition}
\newtheorem{definition}[theorem]{Definition}
\newtheorem{example}[theorem]{Example}
\newtheorem{xca}[theorem]{Exercise}
\theoremstyle{remark}
\newtheorem{remark}[theorem]{Remark}
\[email protected]{@math.fu-berlin.de}
\definecolor{alert}{rgb}{0.8,0,0.3}
\newcommand{\alert}[1]{%
\marginpar{%
\ifodd\value{page} \raggedright \else \raggedleft \fi
\footnotesize{\textcolor{alert}{#1}}
}
}
\begin{document}
\title[Local estimates and uniqueness for the bi-harmonic heat flow]{Some local estimates and a uniqueness result for the entire biharmonic heat equation}
\author{Miles Simon}
\address{Miles Simon:
Institut f\"ur Analysis und Numerik (IAN), Universit\"at Magdeburg, Universit\"atsplatz 2, 39106 Magdeburg, Germany}
\email{miles point simon at ovgu point de}
\author{Glen Wheeler}
\address{Glen Wheeler:
Institut f\"ur Analysis und Numerik (IAN), Universit\"at Magdeburg, Universit\"atsplatz 2, 39106 Magdeburg, Germany}
\curraddr{Institute for Mathematics and its Applications, University of Wollongong, Wollongong 2522, Australia}
\email{<EMAIL>}
\subjclass[2010]{35B45, 35K25, 35K35}
\date{February 2013}
\dedicatory{}
\keywords{fourth-order parabolic partial differential equations, heat flow, local estimates, uniqueness}
\begin{abstract}
We consider smooth solutions to the biharmonic heat equation on
$\mathbb R^n\times [0,T] $ for
which the square of the
Laplacian at time $t$ is globally bounded from above by $k_0/t$ for some $k_0$ in
$\mathbb R^+$, for all $t \in [0,T]$.
We prove local, in space and time, estimates for such solutions. We
explain how these estimates imply uniqueness of smooth solutions in
this class.
\end{abstract}
\maketitle
\section{Introduction}
In this paper we prove local in space and time estimates for solutions $u:\mathbb R^n\times[0,T]\to\mathbb R$ of the biharmonic heat
flow,
\begin{align}
\frac{\partial }{\partial t} u = -\Delta^2 u, \label{bi}
\end{align}
assuming that we have some global in time control on how the solution
behaves as $t\searrow 0$.
This control takes the form
\begin{equation}
\label{EQgc}
t |\Delta u|^2(x,t) \leq k_0 < \infty
\end{equation}
for $t \in [0,T]$ for all $x \in \mathbb R^n$ for some fixed $k_0 \in \mathbb R^+$.
This means that it is possible for $ |\Delta u|^2(x,t)$ to approach infinity as $t \searrow 0$, but if it does so, then
we have some control over the rate at which this occurs.
Here $\Delta u$ refers to the spatial Lapacian of $u$, $\Delta u (x,t) =
\sum_{i=1}^n \nabla_i \nabla_i u(x,t)$ where $\nabla_i u(x,t)$ is the partial
derivative of $u$ with respect to $x_i$.
The growth condition \eqref{EQgc} is natural in the following senses.
It is scale invariant: see the
explanation of the scale invariance of $b$ just after the
the definition of \eqref{assum1} in Section \ref{STblowup}.
This behaviour also does occur in an asymptotic sense. That is, it is
possible to construct a solution $ u \in C^{\infty}( \mathbb R^n \times (0,T))$ and to find points $x(t) \in
\mathbb R^n $ for all $t>0$ such that
$|\Delta u|^2 (x(t),t) = \frac{k_0}{t}$ for some fixed $k_0 >0$ for
all $t>0$.
We also find points $y(t) \in \mathbb R^n$ for all $t>0$ such that
$(\Delta^2 u) (y(t),t) = \frac{k_0}{t}$ for some fixed $k_0 \neq 0$ for
all $t>0$.
That is, the speed of $u$ is not integrable in time.
See the example in Section \ref{STexample}.
Our first result is the following local estimate.
\begin{theorem}\label{mainintro1}
Let $u:\mathbb R^n \times [0,T] \to \mathbb R$, $T < \infty,$ be a smooth solution to \eqref{bi}
that satisfies
\begin{align}
| \Delta u |^2(x,t) \leq \frac{k_0}{t} \label{introA1}
\end{align}
for some $k_0 \in \mathbb R$, for all $t \in [0,T]$, all $x \in \mathbb R^n$ and
\begin{align*}
\sup_{B_1(0)} \sum_{i=0}^{2n+2} |\nabla^i u_0|^2 \leq k_1< \infty,
\end{align*}
for some $k_1 \in \mathbb R$, where $u_0(\cdot) := u(\cdot,0)$.
Then there exists an $N = N(n,k_0,k_1) >0$ such that
\begin{align*}
|\Delta u|^2(x,t) \leq \frac{N}{(1-|x|)^4}
\end{align*}
for all $x$, $t$ which satisfy $x \in \overline B_1(0)$, $\,(1-|x|)^4 \geq
Nt$, and $\,t \leq \frac1N$ , $t \leq T$.
\end{theorem}
For $i \in \mathbb N$ in the above and all that follows, $\nabla^i u(x,t)$ refers to
the full $i$-th spatial derivative, and $|\nabla^i u(x,t)| $ the standard norm
thereof. For example: $\nabla^2 u(x,t) = (\nabla_i \nabla_j u(x,t))_{i,j \in \{1,
\ldots, n\} }$ and $|\nabla^2 u(x,t)|^2 = \sum_{i,j = 1}^n |\nabla_i \nabla_j
u(x,t)|^2$. The operator $\Delta^k$ is defined by $\Delta^k = (\Delta)^k$.
If we have control on other derivatives as $t \searrow 0$ then we obtain
results on higher regularity.
\begin{theorem}\label{mainintro2}
Let $s \in \mathbb N$, $s \geq 2$ and $u:\mathbb R^n \times [0,T] \to \mathbb R$, $T < \infty$ be a smooth solution to \eqref{bi}
which satisfies
\begin{equation}
\label{introA1s}
\left( |\nabla^s u| + |\nabla^{s-1} u|^{s/(s-1)} + \ldots + | \nabla u |^{s}
\right)(x,t)
\leq \frac{k_0}{ t^{s/4}}
\end{equation}
for some $k_0 \in \mathbb R$, for all $t \in [0,T]$, all $x \in \mathbb R^n$ and
\begin{align*}
\sup_{B_1(0)} \sum_{i=0}^{2n+s+1} |\nabla^i u_0|^2 \leq k_1< \infty,
\end{align*}
for some $k_1 \in \mathbb R$.
Then there exists an $N = N(n,k_0,k_1,s) >0$ such that
\begin{align*}
\left( |\nabla^s u| + |\nabla^{s-1} u|^{s/(s-1)} + \ldots + | \nabla u |^{s}
\right)(x,t)
\leq \frac{N}{(1-|x|)^{s}}
\end{align*}
for all $x$, $t$ which satisfy $x \in \overline B_1(0)$, $\,(1-|x|)^4 \geq
Nt$, and $\,t \leq \frac1N$, $t \leq T$.
\end{theorem}
In Section \ref{Tych} we construct an example of a solution to \eqref{bi} which
starts off indentically equal to zero, becomes immediately non-zero and is
smooth in space and time. Solutions of this type for the heat equation are
known to exist and were constructed by Tychonoff, see \cite{Tych}. In
particular, smooth solutions are not uniquely determined by their initial
values: $u(\cdot,\cdot) = 0$ is also a solution. If however we consider
smooth solutions which satisfy \eqref{introA1} then the solution is uniquely
determined by it's initial value, as we show in Section \ref{uniqueness}. The
theorem we prove is:
\begin{theorem}\label{uniquenessintro}
Let $u,v:\mathbb R^n \times [0,T] \to \mathbb R$, $T < \infty$ be smooth solutions to \eqref{bi}
which satisfy
\begin{align*}
|\Delta v|^2(x,t) + |\Delta u |^2(x,t) \leq \frac{k_0}{ t}
\end{align*}
for some $k_0 \in \mathbb R$,
for all $t \in [0,T]$ and all $x \in \mathbb R^n$ and
\begin{align*}
u_0(\cdot) = v_0(\cdot).
\end{align*}
Then $u \equiv v$.
\end{theorem}
The uniqueness problem for the classical heat flow has a rich history.
In the setting where one assumes the solution is non-negative, D. Widder
established uniqueness for the heat flow on $\mathbb R$ for solutions whose
initial value is zero, see Theorem 5 of \cite{W}.
His method relied upon a specific integral representation of the
solution, Theorem 4 of \cite{W}, which
is valid for non-negative solutions.
This proved to be readily generalisable and so influential as to be given its
own name: a uniquness theorem is of Widder-type if the only hypotheses are on
the geometry of the ambient space and that the solution be non-negative.
For example, Aronson \cite{Ar} proved that non-negative solutions to second
order linear equations of divergence form (the coefficents of the
operator being sufficiently regular) in $\mathbb R^n$ are uniquely determined by
their initial data: see Section 5 of that paper.
Here we take an approach reminiscent of \cite{Si1} that
complements the existing literature. Our assumptions for Theorem
\ref{uniquenessintro} are global but we do not require any non-negativity of
the solution.
Although the flow \eqref{bi} is higher-order, we are able to obtain our
estimates using pointwise assumptions, as opposed to integral conditions.
The paper is organised as follows.
Section \ref{STblowup} contains the proof of Theorem \ref{mainintro1} and
Theorem \ref{mainintro2}.
These proofs require some energy estimates for solutions to \eqref{bi}, which
is the subject of Section \ref{STaprioriestimates}.
Section \ref{STuniqueness} contains the proof of Theorem \ref{uniquenessintro},
and Section \ref{Tych} contains full details on the Tychonoff-type solutions
discussed above.
We present in Section \ref{STexample} details on the construction of the example
mentioned above which shows that control of the form \eqref{EQgc} is
natural.
We also show that the solution in this example has a speed which is not integrable.
Some of the estimates from Section \ref{STblowup} and Section
\ref{STaprioriestimates} rely on interpolation inequalities which are not
readily available in the current literature. These interpolation inequalities
are proved in the Appendix.
\section{A Blowup argument}
\label{STblowup}
Let $u: \mathbb R^n \times [0,T] \to \mathbb R$ be a
smooth solution to
\eqref{bi}.
We consider the scale invariant quantity
$e(u): B_1(0) \times [0,T] \to \mathbb R$ defined by
\begin{align*}
e(u)(x,t):= d^4(x) |\Delta u|^2(x,t),
\end{align*}
where $d(x):= (1- |x|)$ is the distance from the boundary
$\partial B_1(0)= \{ x \in \mathbb R^n \,|\, |x| = 1 \}$ to the point $x$ in
$B_1(0)$.
Note that $e = 0$ on $\partial B_1(0)$.
The function $e$ is scale invariant in the following sense:
If we define $\tilde u( \tilde x, \tilde t):= u(c \tilde x, c^4
\tilde t)-c_0,$ where $c_0$ is an arbitrary constant in $\mathbb R$,
then $\tilde u: \mathbb R^n \times [0,\tilde T] \to \mathbb R$ is still a
smooth solution to \eqref{bi} and the quantity $\tilde e( \tilde x, \tilde t)
:= \tilde e(\tilde u)(\tilde x,\tilde t) $
which is defined by
\begin{align*}
\tilde e(\tilde u)(\tilde x,\tilde t):= {\tilde d}^4(\tilde x)
|\Delta \tilde u|^2(\tilde x,\tilde t) ,
\end{align*}
satisfies
\begin{align*}
\tilde e(\tilde x,\tilde t) := e(x,t),
\end{align*}
where here
$ x:= c \tilde x$, $ t:= c^4
\tilde t$ , $\tilde T = \frac{ T}{c^4} $, and ${\tilde d}(\tilde x) := ( \frac 1 c -
| \tilde x|)$ is the distance from $x \in B_{1/c}(0)$ to $ \partial B_{1/c}(0)$.
The scale invariance of $e$ can be seen as follows:
\begin{align*}
(\nabla \tilde u) (\tilde x, \tilde t) &= c (\nabla u) (x, t)\,,\ &&\mbox{and hence} \
(\nabla^k \tilde u ) (\tilde x, \tilde t) = c^k ( \nabla^k u ) (x, t)\,,
\\
\left(\frac{\partial{}}{{\partial t}} \tilde u \right)(\tilde x, \tilde t) &= c^4 \Big(\frac{\partial }{\partial t} u\Big) ( x, t)\,,
\ &&\hskip-1cm\mbox{and hence}\ \left(\Big(\frac {\partial } {\partial t}\Big)^k
\tilde u\right) (\tilde x, \tilde t) = c^{4k} \left( \Big(\frac{\partial }{\partial t} \Big)^k u\right) ( x,
t)\,,\\
\tilde d( \tilde x) &= \Big( \frac 1 c - | \tilde x|\Big)
&&\\&= \frac 1 c ( 1 - | c \tilde
x|)&&\\ &= \frac 1 c ( 1 - |x|) = \frac 1 c d(x) \,,
&&\mbox{and hence}\
{\tilde d}^4( \tilde x) = \frac 1 {c^4} d^4(x)\,,
\end{align*}
where here $(\frac{\partial }{\partial t} )^k$ refers to $k$ time derivatives, and
$\nabla^ku$ refers to $k$ spatial derivatives, and we are assuming that
$k \in \mathbb N$ ($k \neq0$).
Therefore
\begin{align*}
|\Delta \tilde u|^2(\tilde x, \tilde t){\tilde d}^4(\tilde x) = |c^2 \Delta u|^2 (x,t)
\frac 1 {c^4} d^4(x) = |\Delta u|^2 (x,t) d^4(x).
\end{align*}
and $e(x,t) =
\tilde e(\tilde x, \tilde t)$ as claimed.
Note that
\begin{align*}
x \in B_1(0)\,, \ d^4(x) \geq Nt\quad \Longleftrightarrow \quad
\tilde x \in B_{1/c}(0)\,, \ {\tilde d}^4(\tilde x) \geq N \tilde t
\end{align*}
in view of the definitions of the terms involved.
In the following, we will assume that
\begin{equation}
\tag{\rm A1} b(u)(x,t) := t |\Delta u|^2(x,t) \leq k_0 < \infty
\ \mbox{ for all } x \in \mathbb R^n\,, t \in [0,T], \label{assum1}
\end{equation}
for some fixed $k_0 \in \mathbb R^+$.
That is, the quantity \begin{equation}
Q(x,t):= Q(u)(x,t) := |\Delta u|^2(x,t) \label{Qdefn}
\end{equation}
may approach infinity
as $t \searrow 0$, but it is only allowed to do so at a controlled,
but non-integrable rate.
Note that we have $e(x,t) = d^4(x) Q(x,t) $ and $b(x,t) = t Q(x,t)$.
The function $b(x,t)$ is also scale invariant in the following sense:
If we define $\tilde u$, $\tilde x$ and $\tilde t $ as above, then
$\tilde b(\tilde x,\tilde t) := b(\tilde u)(\tilde x, \tilde t) = b(x,t)$
and hence
$\tilde b(\tilde x,\tilde t)\leq k_0 < \infty$ for all $\tilde x \in
\mathbb R^n$ and all $\tilde t \in [0,\tilde T]$.
The scale invariance of $b$ may be verified with an argument
similar to the one
we used above to show that $e$ is scale invariant.
We are interested in the local behaviour of solutions $u$ to \eqref{bi} which satisfy \eqref{assum1}.
In particular, if at time zero $u_0 = u(\cdot,0)$ satisfies
\begin{equation}
\tag{A2} \sup_{B_1(0)} \sum_{i=0}^{2n+2} |\nabla^i u_0|^2 \leq k_1< \infty,
\label{assum2}
\end{equation}
for some fixed $k_1 \in \mathbb R^+$, then we show that the solution satisfies estimates on a smaller ball
for a short well-defined time interval.
The following theorem is Theorem \ref{mainintro1} of the introduction.
\begin{theorem}\label{main1}
Let $u:\mathbb R^n \times [0,T] \to \mathbb R$, $T < \infty$ be a smooth solution to \eqref{bi}
which satisfies assumptions \eqref{assum1} and \eqref{assum2}.
Then there exists an $N = N(n,k_0,k_1) >0$ such that
\begin{align*}
e(x,t) \leq N
\end{align*}
for all $x$, $t$ which satisfy $x \in \overline B_1(0)$, $\,d^4(x)
\geq Nt$, and $\,t \leq \frac1N$, $t \leq T$.
\end{theorem}
For our theorem on higher order regularity, we modify the quantities
above.
Let $s \in \mathbb N$, $s \geq 2$ be given and fixed, and define
\begin{align*}
& Q_s(u)(x,t) = ( |\nabla^s u| + |\nabla^{s-1} u|^{s/(s-1)} +
\ldots + | \nabla u |^{s} )^{4/s}(x,t) \\
& e_s(u)(x,t) = d^4(x) Q_s(u)(x,t) \\
& b_s(u)(x,t) = t Q_s(u)(x,t) .
\end{align*}
These quantities are scale invariant in the sense explained above:
for $\tilde u$, $\tilde T$, $\tilde t$, $\tilde x$ and $\tilde d$ defined as above, and
$\tilde Q_s(\tilde x , \tilde t) := Q_s(\tilde u)(\tilde x, \tilde t)$ we have
\begin{align*}
\tilde e_s(\tilde x,\tilde t) &:= {\tilde d}^4(\tilde x) \tilde Q_s (\tilde x,\tilde t) = d^4(x) Q_s(u)(x,t)\,,\quad\text{and}\\
\tilde b_s(\tilde x, \tilde t) &:= \tilde t \tilde Q_s (\tilde x,\tilde t) = t Q_s (x, t) .
\end{align*}
For this set-up we require
\begin{equation}
\tag{${\rm A}_{\rm s}1$}
\label{assumt1}
b_s(u)(x,t) \leq k_0 < \infty
\end{equation}
for all $x \in \mathbb R^n$ $t\in [0,T]$, and for some fixed $k_0\in\mathbb R^+$, $s \geq 2$,
$s \in \mathbb N$; and
\begin{align}
\tag{${\rm A}_{\rm s}2$}
\label{assumt2}
\sup_{B_1(0)} \sum_{i=0}^{2n+s+1} |\nabla^i u_0|^2 \leq k_1< \infty,
\end{align}
for some fixed $k_1 \in \mathbb R^+$.
In this context we obtain the following variant of Theorem \ref{main1} above.
\begin{theorem}\label{main2}
Let $u:\mathbb R^n \times [0,T] \to \mathbb R$, $T < \infty$ be a smooth solution to \eqref{bi}
which satisfies assumptions ($A_s$1) for some $s \in N, s\geq 2$ and ($A_s$2).
Then there exists an $N = N(n,k_0,k_1,s) >0$ such that
\begin{align}
e_s(x,t) \leq N
\,,\label{conclusion2}
\end{align}
for all $x$, $t$ which satisfy $x \in \overline B_1(0)$, $d^4(x) \geq Nt$, $t
\leq \frac1N$, $t \leq T$. \end{theorem}
Note that this theorem is equivalent to
Theorem \ref{mainintro2} of the introduction.
Under the same assumptions as Theorem \ref{main2}, but with the condition that $s \geq 4$ , we also obtain a local supremum
bound for $|u|$:
\begin{corollary}\label{main3}
Assume everything is as in Theorem \ref{main2} but that $s \geq 4$. Then we also have
\begin{align}
|u(x,t)| \leq \sqrt{ k_1 } + 1 \label{conclusion3}
\end{align}
for all $x$, $t$ which satisfy $x \in \overline B_1(0)$, $d^4(x) \geq Nt$, $t
\leq \frac1N$, where $N$ is as in the conclusion of Theorem \ref{main2}
above.
\end{corollary}
\begin{proof}[Proof of Corollary \ref{main3}.]
Let $(x,t_0)$ be a point which satisfies $d^4(x) \geq Nt_0$ and $t_0 \leq
\frac 1 N$.
Then $d^4(x) \geq Nt $ and $ t \leq \frac 1 N$ for all $t \leq t_0$.
Hence, taking $s=4$ in Theorem \ref{main2} we see that $|\Delta^2 u |(x,t)| \leq
\frac{N}{d^4(x)}$ for all $t \leq t_0$ and that $|u(x,0)| \leq \sqrt{k_1}$ in view
of \eqref{assumt2}.
The evolution equation for $u(x,t)$ is
$\frac{\partial }{\partial t} u(x,t) = -\Delta^2 u (x,t)$. Integrating this from $0$ to $t_0$
and using the two estimates which we just derived, we see that
$ |u(x,t_0)| \leq t_0 \frac{N}{d^4(x)} +\sqrt{ k_1 }$. Using that $d^4(x) \geq
N t_0$
we obtain the result.
\end{proof}
Now we prove Theorem \ref{main1}.
\begin{proof}[Proof of Theorem \ref{main1}.]
Define $d$ resp. $e$ to
be zero on $\mathbb R^n \backslash B_1(0)$ resp. $(\mathbb R^n \backslash B_1(0))
\times [0,T]$. Then $e$ is continuous on $\mathbb R^n \times [0,T]$.
Assume that the conclusion of the theorem is false, and let $N \in
\mathbb N$.
Note that $e(x,0) \leq k_0$ where $k_0$ is the constant appearing in
\eqref{assum2}. Without loss of generality $N >k_0$.
The set of $x \in \overline{B_1(0)} $, $ t\in [0,T]$ for which
$1 \geq d^4(x) \geq Nt$ and $t
\leq \frac 1 N$ is a compact set in $\mathbb R^n \times [0,T]$ which we
denote by $K$.
By compactness of $K$ and continuity of $e$ and the fact that $e(x,0)
\leq k_0 < N$ for all $x \in \overline{B_1(0)}$ , there must be a first time $t_0 \in (0,\frac 1 N]$ and (at least) one point
$x_0 \in B_1(0)$ such that $e(x_0,t_0) = N$. That is:
$e(x,t) < N$ for all $(x,t) \in K$ with $ t < t_0$, and
$e(x_0,t_0) = N$ for some point $(x_0,t_0) \in K$. Clearly we have
$d(x_0) >0$ for such a point, that is, $x_0 \in B_1(0)$, since $e(x_0,t_0)
> 0$.
Rescale the solution $u$ to $\tilde u( \tilde x, \tilde t):=
u(c \tilde x, c^4 \tilde t) -c_0$, where $c_0:= u(c \tilde x_0,0)$, and $c>0$
is chosen so that ${\tilde d}^4(\tilde
x_0) = N$.
It is possible to choose $c$ in this way: $\tilde d(\tilde x) = \frac 1 c d(x)$, so we choose $c^4 =
\frac{d^4(x_0)}{N}$, which is larger than zero since $d(x_0) > 0$ as we
explained above.
Our choice of $c_0 $ guarantees that $\tilde u(\tilde x_0,0) = 0$.
Note for later use that $c^4 = \frac{d^4(x_0)}{N} \leq \frac 1 N\, (\leq
1)$, and $c \searrow 0$ as $ N \to \infty$.
Now
\begin{align*}
N = e(x_0,t_0) = \tilde e(\tilde x_0,\tilde t_0)
= {\tilde d}^4(\tilde x_0) \tilde Q(\tilde
x_0,\tilde t_0) = N \tilde Q(\tilde x_0,\tilde t_0)
\end{align*}
due to scaling, and hence
\begin{align*}
\tilde Q(\tilde x_0,\tilde t_0) = 1.
\end{align*}
Similarly,
\begin{align*}
N \geq e(x,t)
= \tilde e(\tilde x,\tilde t)
= {\tilde d}^4(\tilde x) \tilde Q(\tilde
x,\tilde t)
\end{align*}
for all $(x,t) \in K$ with $t \leq t_0$, implies
\begin{align}
\tilde Q( \tilde x , \tilde t) \leq \frac N { {\tilde d}^4(\tilde x) } \label{mainest}
\end{align}
for all $\tilde x \in B_{1/c}(0)$ with ${\tilde d}^4(\tilde x) \geq \tilde t N$ and
$\tilde t \leq \tilde t_0$.
Note that the inequality \eqref{mainest} is also valid for
all $\tilde x $ with ${\tilde d}^4(\tilde x) \geq \tilde t N$ and
$\tilde t \leq \tilde t_0$, since $\tilde d (\tilde x) = 0$ outside of $
B_{1/c}(0)$ (here we define $\frac{M}{0} = \infty$ for $M>0$).
As in the paper \cite{Si1} we consider two cases:
{\bf Case 1}, where ${\tilde d}^4(\tilde x_0) \geq 2N\tilde t_0$ (which is
equivalent to $\tilde t_0 \leq \frac 1 2$, since ${\tilde d}^4(\tilde x_0) = N$), and
{\bf Case 2}, where ${\tilde d}^4(\tilde x_0) < 2N\tilde t_0$ (which is
equivalent to $1 \geq \tilde t_0 > \frac 1 2$, since $\tilde t_0 N \leq {\tilde d}^4(\tilde x_0) <2N\tilde t_0$ and ${\tilde d}^4(\tilde x_0 ) = N$.
Note that $N \tilde t_0 \leq {\tilde d}^4(\tilde x_0) $ since $(x_0,t_0) \in K$
).
We start with Case 1.
\vspace*{5mm}
\noindent {\bf Case 1}: In this case we have ${\tilde d}^4(\tilde x_0) \geq 2N\tilde t_0$.
For $\tilde y $ with ${\tilde d}^4(\tilde y) \geq \frac N 2$, we obtain
\begin{align}
{\tilde d}^4(\tilde y) \geq \frac N 2 \geq N \tilde t_0 \geq N \tilde t
\label{(*)}
\end{align}
for all $ \tilde t \leq \tilde t_0$.
Hence, we see that
\begin{align}
\tilde Q( \tilde y, \tilde t)
\leq \frac N { {\tilde d}^4(\tilde y) }
\leq 2 \label{mainest2}
\end{align}
for all $\tilde t \leq \tilde t_0$
in view of \eqref{(*)} and \eqref{mainest}.
We also have that ${\tilde d}^4(\tilde x_0) = N \geq \frac N 2$ and so the above
estimate also holds for $\tilde y = \tilde x_0$ and $\tilde t = \tilde t_0$.
We calculate
\begin{align}
\frac N 2
&\leq {\tilde d}^4(\tilde y) = \left( \frac 1 c - |\tilde y| \right)^4
\notag\\
\Longleftrightarrow \quad \left(\frac 1 c - |\tilde y|\right)
&\geq \frac{ N^{ \frac 1 4} }{2^{\frac 1 4} }
\notag\\
\Longleftrightarrow \hspace{3mm}\qquad\quad\ \ |\tilde y|
&\leq - \frac{ N^{\frac 1 4} }{ 2^{\frac 1 4} } + \frac 1 c
\,.
\label{(**)}
\end{align}
Furthermore ${\tilde d}^4 (\tilde x_0) = N$ implies
$|\tilde x_0| = -N^{\frac 1 4} + \frac 1 c$.
Assume that $\tilde y $ is an arbitrary point with
$\tilde y \in B_{ N^{\frac 1 4} / 400 }(\tilde x_0)$.
Then we have
\begin{align}
|\tilde y| \leq | \tilde x _0 | + | \tilde x_0 - \tilde y |
&= -N^{\frac 1 4} + \frac 1 c + | \tilde x_0 - \tilde y | \cr
&\leq -N^{\frac 1 4} + \frac 1 c + \frac {N^{\frac 1 4}}{400} \cr
&\leq \frac 1 c - \frac {N^{\frac 1 4}}{ 2^{\frac 1 4}} \label{didly}
\intertext{and hence, in view of \eqref{(**)}}
{\tilde d}^4(\tilde y) &\geq \frac N 2\,.
\end{align}
Therefore $\tilde y \in B_{ N^{\frac 1 4}/ 400 }(\tilde x_0)$ and $\tilde t\leq \tilde
t_0 \leq \frac 1 2$ implies in Case 1 that
\begin{align*}
\tilde Q( \tilde y, \tilde t) \leq 2\,,\quad\text{and}\quad
\tilde Q( \tilde x_0, \tilde t_0) = 1\,,
\end{align*}
in view of \eqref{mainest2} and the definition of $\tilde x_0$ and $\tilde
t_0$.
\vspace*{5mm}
\noindent {\bf Case 2.}
In this case we have $1\geq \tilde t_0 > \frac 1 2.$
For all $\tilde t \leq \frac 1 2$ and $\tilde y$ with ${\tilde d}^4(\tilde y) \geq \frac N 2$ we
have
\begin{align*}
{\tilde d}^4(\tilde y) \geq \frac N 2 \geq N \tilde t
\end{align*}
and hence
\begin{align}
\tilde Q( \tilde y, \tilde t) & \leq \frac N { {\tilde d}^4(\tilde y)} \leq 2 \label{zwisch1}
\end{align}
in view of \eqref{mainest}.
For $\tilde t_0 \geq \tilde t \geq \frac 1 2$ we have
\begin{align}
\tilde Q( \tilde y, \tilde t) & \leq \frac {k_0}{ \tilde t} \leq 2
k_0 \label{zwisch2},
\end{align}
in view of \eqref{assum1}.
Note that we may assume without loss of generality that $k_0 \geq 1$.
Now we know from \eqref{didly} that
$y \in B_{ N^{\frac 1 4} /400}(\tilde x_0)$ implies that $\tilde
d^4(\tilde y) \geq \frac N 2 $.
Hence, using the inequalities \eqref{zwisch1} and \eqref{zwisch2}, we see that
\begin{equation}
\tilde Q( \tilde y, \tilde t) \leq 2k_0\,,
\quad\text{and}\quad
\tilde Q( \tilde x_0, \tilde t_0) = 1\,,
\label{EQkeyestimate}
\end{equation}
for all $y \in B_{{N^{\frac 1 4}}/{400}}(\tilde x_0)$ and $t \in [0,\tilde t_0]$.
We have shown that in both Case 1 and Case 2 we obtain the estimate \eqref{EQkeyestimate}.
Now we use Corollary \ref{CY2} to obtain a contradiction.
We use $v: B_R(0) \times [0,\tilde t_0] \to \mathbb R$ to denote the rescaled solution $\tilde u:
B_{{N^{\frac 1 4}}/{400}}(\tilde x_0) \times [0,\tilde t_0]
\to \mathbb R$.
That is $v(\cdot,\cdot) = \tilde u(\cdot - \tilde x_0,\cdot)$, $R = N^{\frac 1 4}/400$, $\tilde t_0
\leq 1$. The definition of $\tilde u$ guarantees that $\tilde u(\tilde x_0,0)
=0$, and hence we have $v(0,0)=0$.
Also, using \eqref{assum2} and the fact that $c^4 \leq 1/N$ and $N >1$, we see that
\begin{align}
\sup_{B_{1/c}(0)} \sum_{i=1}^{2n+2} |\nabla^i v|^2(\cdot ,0) & =
\sup_{B_1(x_0)} \sum_{i=1}^{2n+2} c^{2i} |\nabla^i
u|^2(\cdot,0) \cr
& \leq \sup_{B_1(x_0)} \sum_{i=1}^{2n+2} \frac{1}{N^{i/2}} |\nabla^i
u(\cdot ,0)|^2\cr
& \leq \frac{ k_1}{N^{1/2}}.
\end{align}
Hence, combining this estimate with the fact that $v(0,0)=0$ and by
choosing $N$ sufficiently large, we may
assume w.l.o.g. that
\begin{align}
\sup_{B_1(0)} \sum_{i=0}^{2n+2} |\nabla^i v|^2(\tilde x,0) \leq \tilde \epsilon(N) \label{newA2}
\end{align}
where $\tilde \epsilon(N) \to 0$ as $N \to \infty$ ( $k_0,k_1,n$ are fixed in
this argument). Define $2\rho = R = N^{\frac 1
4}/400 \leq {1/c}$ ($c \leq \frac{1}{N^{1/4}}$ as we mentioned
above) and $p = p(n) = n+1$. Then $\rho \to \infty$ as $N \to \infty$.
Corollary \ref{CY2} implies that
\begin{align}
\frac{d}{dt}E^{p}_\eta(v) + E^{p+1}_\eta(v)
\le \frac{C}{\rho^{4p}}\int_{\mathbb R^n}
|\Delta v|^2\gamma^{s-4p }\, \label{step1}
\end{align}
for all $t \leq \tilde t_0$ for all $s> 4p+4$, where $\eta = \gamma^s$, and
$\gamma$ is a cutoff function as in \eqref{EQgamma}, and $C =
C(n,s)$. Choose $s = 4p(n) + 5 = 4n + 9$, so that $C = C(n)$.
We know from the estimate \eqref{EQkeyestimate} that $Q (v) = |\Delta
v|^2 \leq 2k_0$ on $ B_R(0) \times [0,\tilde t_0] $
and
hence, combining this with \eqref{step1} we have
\begin{align}
\frac{d}{dt}E^{p}_\eta(v) + E^{p+1}_\eta(v)
& \leq \frac{C}{\rho^{4p}}\int_{\mathbb R^n}
|\Delta v|^2\gamma^{s-4p} \cr
&\leq \frac{C}{\rho^{4p}}
\int_{B_{2\rho}} 2k_0\cr
& = 2\omega_nk_0C \rho^{n-4p} \label{remember}
\end{align}
which implies that
\begin{align*}
E^{p}_\eta(v)(t) \leq (2\omega_n k_0 C + \omega_n k_1) \rho^{n-4p} = (2\omega_n k_0 C + \omega_n k_1) \rho^{-3n-4}
\end{align*}
for all $t \leq \tilde t_0 \leq 1$
in view of the fact that
\begin{align*}
E^{p}_\eta(v)(0) = \int_{\mathbb R^n} |\Delta^{p} v_0|^2( \tilde x )\gamma^{s}d \tilde x
& \leq \int_{B_{2\rho}(0)} |\Delta^{p} v_0|^2(\tilde x) d \tilde x\cr
& \leq \int_{B_{1/c}(0)} |\Delta^{p} v_0|^2(\tilde x) d \tilde x\cr
& = \int_{B_{1/c} (0) } c^{4p } |\Delta^{p} u_0|^2(c \tilde x) d \tilde x\cr
& = c^{4p - n } \int_{B_1{0}} |\Delta^{p} u_0|^2( x) dx
\cr
& \leq k_1 \omega_n c^{4p - n }\cr
&\leq k_1 \omega_n \rho^{n-4p}
\end{align*}
where we have used assumption \eqref{assum2} again, the definition of
$v$, the scaling properties of the derivatives of $u(cx,0)$,
and the fact that $1/c \geq 2\rho \geq \rho $.
In particular,
\begin{align*}
\int_{B_{1}(0)} |\Delta^{p} v |^2 \leq (\omega_n 2k_0 C + k_1
\omega_n) \rho^{-3n-4}\to 0
\end{align*}
as $\rho \to \infty,$ in view of the fact that $ p = p(n) = n+1$.
We have shown that
\begin{align*}
\int_{B_{1}(0)} |\Delta^{p} v |^2 \leq \epsilon_p(k_0,k_1,n,\rho)
\end{align*}
for all $t\leq \tilde t_0 \leq 1$ where $ \epsilon_p(k_0,k_1,n,\rho) \to 0$ as
$\rho \to \infty$, that is, as $N \to \infty$.
We can similarly show that
\begin{align*}
\int_{B_{1}(0)} |\Delta^{p-1} v |^2 \leq \epsilon_{p-1}(k_0,k_1,n,\rho) .
\end{align*}
We also have
\begin{align*}
\frac{d}{dt} \int_{B_1(0)} |\Delta^{p-2} v|^2 & = 2 \int_{B_1(0)}
(\Delta^{p-2} v)(\Delta ^p v)\cr
& \leq \int_{B_1(0)}
|\Delta^{p-2} v|^2 + \int_{B_1(0)} |\Delta^p v|^2\cr
&\leq
\int_{B_1(0)} |\Delta^{p-2} v|^2 + \epsilon_p
\end{align*}
in view of Young's inequality, and the estimate just shown,
and hence, after integrating in time from $0$ to $\tilde t_0 \leq 1$
we see that
\begin{align}
\int_{B_1(0)} |\Delta^{p-2} v|^2 \leq \epsilon_{p-2}(N) \label{inttime}
\end{align}
for all $t\in [0,\tilde t_0 \leq 1] $ with $\epsilon_{p-2}(N) \to 0$ as $N \to \infty$: we leave out dependence on
$k_1,k_0,n$ since these variables are fixed.
More explictly: $f(t):= e^{-t} (\int_{B_1(0)} |\Delta^{p-2} v|^2)(t) -
2t \epsilon_p $ satisfies $\frac{df}{dt}(t) \leq 0$ for all $0\leq t \leq
\tilde t_0$
and $f(0) = (\int_{B_1(0)}
|\Delta^{p-2} v|^2)(0) \leq \tilde \epsilon(N) \to 0 $ as $N \to \infty$ in
view of \eqref{newA2}, and so, integrating $f$ from $0$ to $t_0$, we
see that the estimate \eqref{inttime} is true.
Continuing in this way, we get, for $N$ sufficiently large:
\begin{equation}
\int_{B_1(0)} |\Delta^{l} v|^2 \leq \epsilon_{l}(N) \label{lapsmall}
\end{equation}
for $l = p,p-2,p-4, \ldots, 1$ (or $0$).
Starting with $p-1$ instead of $p$, we similarly get
\begin{equation}
\int_{B_1(0)} |\Delta^{l} v|^2 \leq \epsilon_{l}(N) \label{lapsmall2}
\end{equation}
for $l = p-1,p-3,p-5, \ldots, 1$ (or $0$), where $\epsilon_{l}(N) \to 0$ as $N \to
\infty$.
That is \begin{equation}
\int_{B_1(0)} |\Delta^{l} v|^2 \leq \epsilon_{l}(N) \label{lapsmall3}
\end{equation}
for $l \in \{0, \ldots, p\}$, where $\epsilon_{l}(N) \to 0$ as $N \to
\infty$.
Using the $L^2$ estimates, Lemma \ref{L2est} from the Appendix, and
the estimate \eqref{lapsmall3} we see that
\begin{equation}
\int_{B_{1/2}(0)} |\nabla^l v|^2 \leq \hat \epsilon (N)
\end{equation}
for all $0\leq l \leq 2p= 2n +2$, where $\hat \epsilon(N) \to 0$ as $N \to
\infty$ (choose $\sigma = \sigma(n) = \frac{1}{4p(n)}$, so that $1- 2p \sigma
= 1/2$).
Applying the Sobolev-Morrey inequality \cite[Theorem 6, Section 5.6.3]{Evans}, with
$p$, $k$ there equal to 2, and $2n+2$ respectively, we see that
\[
\tilde Q(0,t_0) = |\Delta v|^2(0,t_0) \leq C(n) \Big(\sum_{l=0}^{2n+2}
\int_{B_{1/2}(0)} |\nabla^l v|^2(\cdot,t_0)\Big)^{\frac 1 2} \leq
C(2n +3)^{\frac 1 2} (\hat \epsilon)^{\frac 1 2}
\]
and hence
\[
\tilde Q(0,t_0) \to 0
\]
as $N \to \infty$. This contradicts the fact that $\tilde Q(0,t_0) = 1$.
\end{proof}
The proof of Theorem \ref{main2} is essentially the same as the proof of Theorem \ref{main1}.
\begin{proof}[Proof of Theorem \ref{main2}.]
Replace $Q(x,t)$ by $Q_s(x,t)$, $e(x,t)$ by $e_s(x,t)$, $b(x,t)$
by $b_s(x,t)$, and $Q(u)$ by $Q_s(u)$ and repeat the above proof. At the point where
$|\Delta v|^2 = Q (v) \leq 2k_0$ on $ B_R(0) \times [0,\tilde t_0] $ is used in the
inequality \eqref{remember}, use instead the
fact that
$|\Delta v|^2 \leq |\nabla^2 v|^2 \leq Q_s (v) \leq 2k_0$.
Also choose $p(n)= n+(s/2)$ or $p= n + \frac{(s+1)}{2}$ in the proof: whichever is an integer.
The last part of the proof, where Morrey's embedding Theorem is used, has to be
slightly modified:
$Q_s (v)(0,t_0) = 1$ implies that $|\nabla^r v|(0,t_0) \geq
\delta(s)>0$ for some $r \in \{ 1, \ldots, s\}$ for some small $\delta(s) >0$:
otherwise
the sum of the terms appearing in $Q_s(v)(0,t_0)$ would be less
than one.
Applying the Sobolev-Morrey inequality \cite[Theorem 6, Section 5.6.3]{Evans} with
$p,k$ there equal to 2, $2n+s$ respectively, we see that
\begin{align*}
0 &< \delta^2(s) \leq |\nabla^r v|^2(0,t_0)
\\
&\leq C(n,s) \Big( \sum_{l=0}^{2n+s} \int_{B_{1/2}(0)}|\nabla^l v|^2(\cdot,t_0) \Big)^{1/2}
\\&\leq C (2n+1+s)^{\frac 1 2} (\hat \epsilon (N))^{1/2},
\end{align*}
which leads to a contradiction if $N$ is chosen large enough, since $\hat
\epsilon(N) \to 0$ as $ N \to \infty$.
\end{proof}
\section{A priori Energy Estimates}
\label{STaprioriestimates}
In this section we shall prove some estimates for the weighted energies
\begin{equation}
E^k_\eta(u) = \int_{\mathbb R^n} |\Delta^ku|^2\,\eta,\quad \eta\in C^\infty_{loc}(\mathbb R^n),\quad \text{supp }\eta \subset\subset \mathbb R^n,
\label{EQenrgy}
\end{equation}
where $k\in\mathbb N_0$ and $u:\mathbb R^n\times[0,T] \rightarrow\mathbb R$ is a smooth solution to \eqref{bi}.
In the above equation and in that which follows all integrals are with respect to Lebesgue measure.
Note that $E^k_\eta$ are all finite for
any $k\in\mathbb N_0$ and $t\in[0,T]$ since $u: \mathbb R^n \times [0,T] \to \mathbb R$ is smooth.
The purpose of the a priori estimates in this section is to quantify how global quantities such as the various
Sobolev norms of the solution behave along the flow \eqref{bi}.
\begin{lemma}
\label{LMlm1}
Let $u:\mathbb R^n\times[0,T]\rightarrow\mathbb R$ be a smooth solution to \eqref{bi}.
For all $t\in[0,T]$,
\begin{align}
\frac{d}{dt}&E^k_\eta(u) + 2E^{k+1}_\eta(u)
\notag\\
&=
- 2\int_{\mathbb R^n} (\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\eta)\,
+ 2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)\,.
\label{EQlm1}
\end{align}
for all $k \in \mathbb N_0$.
\end{lemma}
\begin{proof
Differentiating,
\begin{align*}
\frac{d}{dt}E^k_\eta(u)
&= 2\int_{\mathbb R^n} (\Delta^ku)(-\Delta^{k+2}u)\,\eta\,
\\
&= 2\int_{\mathbb R^n} (\nabla_i\Delta^ku)(\nabla_i\Delta^{k+1}u)\,\eta\,
+ 2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)\,
\\
&= -2\int_{\mathbb R^n} |\Delta^{k+1}u|^2\,\eta\,
- 2\int_{\mathbb R^n} (\nabla_i\Delta^ku)(\nabla_i\eta) (\Delta^{k+1}u)\,
\\&\qquad
+ 2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)\,
\\
&= -2E^{k+1}_\eta(u)
\\&\qquad
- 2\int_{\mathbb R^n} (\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\eta)\,
+ 2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)\,.
\end{align*}
Rearranging gives the lemma.
\end{proof
We now specialise by setting $\eta = \gamma^s$, $s>0$ to be chosen, and $\gamma \in
C^\infty_{\text{loc}}(\mathbb R^n)$ satisfying
\begin{equation}
\label{EQgamma}
\tag{\ensuremath{\gamma}}
\chi_{B_\rho(0)} \le \gamma \le \chi_{B_{2\rho}(0)},\quad \rho>0,
\quad\text{and}\quad
|\nabla\gamma| \le \frac{c_{\gamma}}\rho,\quad|\nabla^2\gamma|\le \frac{c_{\gamma}}{\rho^2}\,,
\end{equation}
where $c_{\gamma}\ge1$ is an absolute constant depending only on $n$.
\newcommand{c_{\gamma}}{c_{\gamma}}
In the following proofs we make extensive use of the elementary inequality
\begin{equation}
\label{EQcauchy}
ab \le \varepsilon a^2 + \frac1{4\varepsilon}b^2
\end{equation}
for $a$, $b$ real numbers and $\varepsilon>0$.
\begin{lemma}
\label{LMlm2}
Let $u\in C^\infty_{\text{loc}}(\mathbb R^n)$.
Suppose $\eta=\gamma^s$, $s>8$, and $\gamma$, $c_{\gamma}$ are as in \eqref{EQgamma}.
Then for any $\varepsilon>0$ and for all $k \in \mathbb N$ we have
\begin{align*}
- 2\int_{\mathbb R^n} (\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\eta)\,
&\le \varepsilon E_\eta^{k+1}(u)
+ \frac{c_{1}(\varepsilon,s,n) }{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,,
\end{align*}
where $c_{1}(\varepsilon,s,n) < \infty $ is a constant depending on $
\varepsilon,s$ and $n$.
\end{lemma}
\begin{proof
Throughout the proof $\delta_i$ denote positive parameters to be chosen.
Using \eqref{EQcauchy} and \eqref{EQgamma},
\begin{align}
- 2\int_{\mathbb R^n} (\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\eta)\,
&=
- 2s\int_{\mathbb R^n} (\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\gamma)\gamma^{s-1}
\notag\\
&\le \delta_1 E_\eta^{k+1}(u)
+ \frac{c_{\gamma}^2s^2}{\delta_1\rho^2} \int_{\mathbb R^n} |\nabla\Delta^ku|^2\gamma^{s-2}\,.
\label{EQlm2begeq1}
\end{align}
Now,
\begin{align*}
\int_{\mathbb R^n}& |\nabla\Delta^ku|^2\gamma^{s-2}
\\
&=
-\int_{\mathbb R^n} (\Delta^ku)(\Delta^{k+1}u)\gamma^{s-2}
-(s-2)\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k}u)(\nabla_i\gamma)\gamma^{s-3}
\\
&\le
\rho^2\delta_2\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac1{4\delta_2\rho^2}\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\\
&\qquad
+ \frac12 \int_{\mathbb R^n} |\nabla\Delta^ku|^2\gamma^{s-2}
+ \frac{c_{\gamma}^2(s-2)^2}{2\rho^2}\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}\\
&= \rho^2\delta_2\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac12 \int_{\mathbb R^n} |\nabla\Delta^ku|^2\gamma^{s-2}\\
&\qquad
\frac{1}{2\rho^2}(\frac{1}{2\delta_2} + c_{\gamma}^2(s-2)^2)\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\end{align*}
Absorbing the second term on the right into the left we obtain
\begin{align}
\notag
\int_{\mathbb R^n} &|\nabla\Delta^ku|^2\gamma^{s-2}
\\
&\le
2\rho^2\delta_2\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac1{\rho^2}\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)
\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}\,.
\label{EQlm2mideq1}
\end{align}
We now need an interpolation inequality.
From Lemma \ref{LMinterp1} we know that
\begin{align*}
\frac1{\rho^2}&\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\\
&\leq \delta_3\rho^2\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac{c_{\delta_3}}{\rho^6}
\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\,,
\end{align*}
where $c_{\delta_3} = c_{\delta_3}(\delta_3,s,n) = 2^4\big(\frac1{\delta_3} + 2^9s^4c_{\gamma}^4\big)$.
Using this in \eqref{EQlm2mideq1} we obtain
\begin{align*}
\int_{\mathbb R^n} |\nabla\Delta^ku|^2\gamma^{s-2}
& \leq 2\rho^2 \delta_2 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
\\
& \hskip-1.5cm + \left(\frac{1}{2\delta_2} + c_{\gamma}^2(s-2)^2\right) \Big(
\delta_3\rho^2\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac{c_{\delta_3}}{\rho^6}
\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\Big)
\\
& = \left( 2\rho^2 \delta_2 + \delta_3\rho^2 \left(\frac{1}{2\delta_2} + c_{\gamma}^2(s-2)^2\right)\right)
\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}\\
&\quad + \frac{c_{\delta_3}}{\rho^6} \left(\frac{1}{2\delta_2} + c_{\gamma}^2(s-2)^2\right)
\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\end{align*}
and hence
\begin{align}
\frac1{\rho^2}\int_{\mathbb R^n} |\nabla\Delta^ku|^2\gamma^{s-2}
&\le
\bigg(
\delta_3\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)
+ 2\delta_2\bigg)\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
\notag\\&\quad
+ \frac{c_{\delta_3}}{\rho^8}
\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)
\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,.
\label{EQstar}
\end{align}
Combining \eqref{EQstar} with \eqref{EQlm2begeq1} gives
\begin{align*}
- 2\int_{\mathbb R^n} &(\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\eta)\,
\\&\le
\bigg[\delta_1
+ \frac{c_{\gamma}^2s^2}{\delta_1}
\bigg(
\delta_3\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)
+ 2\delta_2\bigg)\bigg] E_\eta^{k+1}(u)
\\&\qquad
+ \frac{c_{\gamma}^2s^2}{\delta_1}
\frac{c_{\delta_3}}{\rho^8}
\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)
\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\,.
\end{align*}
We choose $\delta_i = \delta_i(s,\epsilon,n) >0$ so that
\begin{equation*}
\bigg[\delta_1
+ \frac{c_{\gamma}^2s^2}{\delta_1}
\bigg(
\delta_3\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)
+ 2\delta_2\bigg)\bigg] \le \varepsilon
\,.
\end{equation*}
This can be achieved by the choice
\begin{align*}
\delta_1 = \frac{\varepsilon}4,\quad
\delta_2 = \frac{\varepsilon^2}{32c_{\gamma}^2s^2},\quad
\delta_3 = \frac{\varepsilon^4}{8c_{\gamma}^4s^2(16s^2 + \varepsilon^2(s-2)^2)}
\end{align*}
for example.
Recalling the definition of $c_{\delta_3} = c(\delta_3,s,n) = 2^4\big(\frac1{\delta_3} + 2^9s^4c_{\gamma}^4\big)$ yields the result.
\end{proof
\begin{lemma}
\label{LMlm3}
Let $u\in C^\infty_{\text{loc}}(\mathbb R^n)$.
Suppose $\eta=\gamma^s$, $s>8$, and $\gamma$, $c_{\gamma}$ are as in \eqref{EQgamma}.
Then for any $\varepsilon>0$ and for all $ k \in \mathbb N$ we have
\begin{align*}
2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)\,
&\le \varepsilon E_\eta^{k+1}(u)
+ \frac{c_{2}(\varepsilon,s,n)}{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,,
\end{align*}
where $c_2 (\varepsilon,s,n) < \infty $ is a constant depending on
$\varepsilon, s , n.$
\end{lemma}
\begin{proof
Again, throughout the proof $\delta_i>0$ denote positive parameters to be chosen.
Integrating by parts,
\begin{align}
2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)\,
&=
- 2\int_{\mathbb R^n} (\nabla_i\Delta^ku)(\Delta^{k+1}u)(\nabla_i\eta)\,
\notag\\&\qquad
- 2\int_{\mathbb R^n} (\Delta^ku)(\Delta^{k+1}u)(\Delta\eta)\,
\label{EQlm3begeq1}
\end{align}
Lemma \ref{LMlm2} deals with the first term:
\begin{equation}
- 2\int_{\mathbb R^n} (\nabla_i\Delta^ku)(\Delta^{k+1}u)(\nabla_i\eta)\,
\le \frac{\varepsilon}{2} E_\eta^{k+1}(u)
+ \frac{c_{1} ( \frac{\varepsilon}{2},s,n)}{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,.
\label{EQlm3begeq2}
\end{equation}
Since $\Delta\eta = s\gamma^{s-1}\Delta\gamma + s(s-1)\gamma^{s-2}|\nabla\gamma|^2$ we have
\begin{align}
- 2\int_{\mathbb R^n}& (\Delta^ku)(\Delta^{k+1}u)(\Delta\eta)\,
\notag\\
\label{EQstarstar}
&\le
2\delta_1\int_{\mathbb R^n} |\Delta^{k+1}u|^2\eta\,
+ \frac{1}{\delta_1\rho^4}\Big(
c_{\gamma}^2s^2 + c_{\gamma}^4s^2(s-1)^2
\Big)\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}\,\, ,
\end{align}
where we used the fact that $\gamma^{s-2}(\cdot) \leq \gamma^{s-4}(\cdot)$,
which is true since $0\leq \gamma(\cdot) \leq 1$ on $\mathbb R^n$.
Lemma \ref{LMinterp1} yields the estimate
\begin{align*}
\frac{1}{\delta_1\rho^4}&\Big(
c_{\gamma}^2s^2 + c_{\gamma}^4s^2(s-1)^2
\Big)\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\\
&\le
\frac{\delta_2}{\delta_1}\Big(
c_{\gamma}^2s^2 + c_{\gamma}^4s^2(s-1)^2
\Big)\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
\\&\quad+
\frac{c_{\delta_2}}{\delta_1\rho^8}\Big(
c_{\gamma}^2s^2 + c_{\gamma}^4s^2(s-1)^2
\Big)\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\end{align*}
where $c_{\delta_2} = c(\delta_2,s,n) =
2^4\big(\frac1{\delta_2} + 2^9s^4c_{\gamma}^4\big)$.
Combining this with \eqref{EQstarstar} we get
\begin{align}
- 2\int_{\mathbb R^n} (\Delta^ku)(\Delta^{k+1}u)(\Delta\eta)\,
&\le
\bigg(2\delta_1 + \frac{\delta_2}{\delta_1}\Big(
c_{\gamma}^2s^2 + c_{\gamma}^4s^2(s-1)^2
\Big)
\bigg)\int_{\mathbb R^n} |\Delta^{k+1}u|^2\eta\,
\notag
\\*&\hskip-1cm
+ \frac{c_{\delta_2}}{\delta_1\rho^{8}}
\left(
c_{\gamma}^2s^2 + c_{\gamma}^4s^2(s-1)^2
\right)\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,.
\label{EQlm3begeq3}
\end{align}
Combining \eqref{EQlm3begeq3} with
\eqref{EQlm3begeq1}-\eqref{EQlm3begeq2} and choosing $\delta_i =
\delta_i(\varepsilon,n,s) >0$ so that
\[
\frac{\delta_2}{\delta_1}\Big(
c_{\gamma}^2s^2 + c_{\gamma}^4s^2(s-1)^2
\Big)
+ 2\delta_1
= \varepsilon/2
\]
yields the result.
One possible choice is
\begin{align*}
\delta_1 = \frac{\varepsilon}{16},\quad
\delta_2 = \frac{\varepsilon^2}{128c_{\gamma}^2\Big( s^2 + c_{\gamma}^2s^2(s-1)^2 \Big)}
\,.
\end{align*}
\end{proof
\begin{corollary}
\label{CY1}
Let $u:\mathbb R^n\times[0,T]\rightarrow\mathbb R$ be a smooth solution to \eqref{bi}.
Suppose $\eta=\gamma^s$, $s>8$, and $\gamma$, $c_\gamma$, are as in
\eqref{EQgamma}, and $k \in \mathbb N$.
Then
\begin{equation*}
\frac{d}{dt}E^k_\eta(u) + \frac32E^{k+1}_\eta(u)
\le \frac{c_3(n,s)}{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}dx\,,
\end{equation*}
where $c_3(n,s)$ is constant depending only on $n$ and $s$.
\end{corollary}
\begin{proof}
We combine Lemmata \ref{LMlm1}--\ref{LMlm3} as follows.
Lemma \ref{LMlm1} states that
\begin{align}
\frac{d}{dt}&E^k_\eta(u) + 2E^{k+1}_\eta(u)
\notag\\
\label{EQcy1eq1}
&=
- 2\int_{\mathbb R^n} (\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\eta)\,
+ 2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)\,.
\end{align}
The two terms on the right hand side are estimated by Lemma \ref{LMlm2} and Lemma \ref{LMlm3} respectively.
Adding together the estimates we find, for any $\varepsilon_1, \varepsilon_2 > 0$,
\begin{align*}
- 2\int_{\mathbb R^n}& (\Delta^{k+1}u)(\nabla_i\Delta^ku)(\nabla_i\eta)
+ 2\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k+1}u)(\nabla_i\eta)
\\
&\le (\varepsilon_1
+ \varepsilon_2) E_\eta^{k+1}(u)
+ \frac{c_1(\varepsilon_1,n,s) + c_{2}(\varepsilon_2,n,s) }{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}.
\end{align*}
In particular choosing $\varepsilon_i = \frac14$ and combining this with \eqref{EQcy1eq1} we have
\begin{align*}
\frac{d}{dt}E^k_\eta(u) + 2E^{k+1}_\eta(u)
\le \frac12
E_\eta^{k+1}(u)
+ \frac{c_3}{2\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,,
\end{align*}
where $c_3$ is a constant depending only on $s$ and $n$.
Absorbing the first term on the right into the left yields the claimed estimate.
\end{proof}
\begin{corollary}
\label{CY2}
Let $u:\mathbb R^n\times[0,T]\rightarrow\mathbb R$ be a smooth solution to \eqref{bi}.
Suppose $k\in\mathbb N $, $\eta=\gamma^s$, $s>4k$, where $\gamma$, $c_{\gamma}$ are as in \eqref{EQgamma}.
Then
\begin{equation*}
\frac{d}{dt}E^k_\eta(u) + E^{k+1}_\eta(u)
\le \frac{c_4(n,s)}{\rho^{4k}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4k}\,,
\end{equation*}
where $c_4(n,s)$ is a constant depending only on $n$ and $s$.
\end{corollary}
\begin{proof}
We first consider the case where $k=1$.
Using Lemma \ref{LMlm1} and integration by parts we find
\begin{align}
\frac{d}{dt}&E^1_\eta(u) + 2E^2_\eta(u)
\notag\\
&=
- 2\int_{\mathbb R^n} (\Delta^2 u)(\nabla_i\Delta u)(\nabla_i\eta)\,
+ 2\int_{\mathbb R^n} (\Delta u)(\nabla_i\Delta^2 u)(\nabla_i\eta)\,
\notag\\
&=
- 4\int_{\mathbb R^n} (\Delta^2 u)(\nabla_i\Delta u)(\nabla_i\eta)\,
- 2\int_{\mathbb R^n} (\Delta u)(\Delta^2 u)(\Delta\eta)\,.
\label{EQzero}
\end{align}
We claim that
\begin{equation}
\label{EQStar}
- 4\int_{\mathbb R^n} (\Delta^2 u)(\nabla_i\Delta u)(\nabla_i\eta)
\le \varepsilon E^2_\eta(u) + \frac{c(\varepsilon,n,s)}{\rho^{4}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4}\,,
\end{equation}
and
\begin{equation}
\label{EQStarstar}
- 2\int_{\mathbb R^n} (\Delta u)(\Delta^2 u)(\Delta\eta)
\le \varepsilon E^2_\eta(u) + \frac{c(\varepsilon,n,s)}{\rho^{4}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4}\,,
\end{equation}
hold.
Given the above estimates, we may conclude the required statement for the case $k=1$ as follows. Choosing $\varepsilon = \frac12$ in each of \eqref{EQStar}, \eqref{EQStarstar} and combining with \eqref{EQzero} we find
\[
\frac{d}{dt}E^1_\eta(u) + 2E^2_\eta(u)
\le E^2_\eta(u) + \frac{c(n,s)}{\rho^{4}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4}\,,
\]
whereupon subtraction of $E^2_\eta(u)$ from both sides yields the desired estimate.
The estimate \eqref{EQStarstar} is \eqref{EQstarstar} with $\delta_1 = \varepsilon/2$ and $k=1$.
It remains to prove the estimate \eqref{EQStar}.
We compute
\begin{equation}
\label{EQstar1}
- 4\int_{\mathbb R^n} (\Delta^2 u)(\nabla_i\Delta u)(\nabla_i\eta)
\le \delta_1 E^2_\eta(u) + \frac{4c_{\gamma}^2s^2}{\delta_1\rho^2}\int_{\mathbb R^n} |\nabla\Delta u|^2\gamma^{s-2}\,.
\end{equation}
Now estimate
\begin{align*}
\int_{\mathbb R^n}& |\nabla\Delta u|^2\gamma^{s-2}
\\
&=
-\int_{\mathbb R^n} (\Delta u)(\Delta^2u)\gamma^{s-2}
-(s-2)\int_{\mathbb R^n} (\Delta u)(\nabla_i\Delta u)(\nabla_i\gamma)\gamma^{s-3}
\\
&\le \rho^2\delta_2\int_{\mathbb R^n} |\Delta^2u|^2\gamma^{s}
+ \frac12 \int_{\mathbb R^n} |\nabla\Delta u|^2\gamma^{s-2}\\
&\qquad
\frac{1}{2\rho^2}\Big(\frac{1}{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4}\,.
\end{align*}
Absorbing the second term on the right into the left we obtain
\begin{align}
\notag
\int_{\mathbb R^n} &|\nabla\Delta u|^2\gamma^{s-2}
\\
&\le
2\rho^2\delta_2\int_{\mathbb R^n} |\Delta^2u|^2\gamma^{s}
+ \frac1{\rho^2}\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)
\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4}\,.
\label{EQstar2}
\end{align}
Combining \eqref{EQstar1} with \eqref{EQstar2} we find
\begin{align*}
- 4&\int_{\mathbb R^n} (\Delta^2 u)(\nabla_i\Delta u)(\nabla_i\eta)
\\
&\le \Big(\delta_1 +
2\rho^2\delta_2\frac{4c_{\gamma}^2s^2}{\delta_1\rho^2}\Big)
E^2_\eta(u)
+ \frac{4c_{\gamma}^2s^2}{\delta_1\rho^2}\bigg(\frac1{\rho^2}\Big(\frac1{2\delta_2} + c_{\gamma}^2(s-2)^2\Big)\bigg) \int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4}\,.
\end{align*}
Choosing $\delta_1>0$ and $\delta_2>0$ such that $\Big(\delta_1 + 2\delta_2\frac{4c_{\gamma}^2s^2}{\delta_1}\Big) \le \varepsilon$ yields \eqref{EQStar}.
Let us now continue by considering the case where $k\ge2$.
In this case we have $k-1 \in \mathbb N$, and $s >4k$ implies $s-4 >4k-4 = 4(k-1)$.
Using these facts and
Corollary \ref{LMinterp2}, we see that
\begin{equation}
\frac{1}{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\le \frac{1}{\rho^4} \int_{\mathbb R^n} |\Delta^{k}u|^2\gamma^{s-4}
+ \frac{c(s,n)}{\rho^{4k}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4k}
\label{EQcy2e1}
\,.
\end{equation}
and, using Corollary \ref{LMinterp2} again,
\begin{equation}
\frac{1}{\rho^4} \int_{\mathbb R^n} |\Delta^{k}u|^2\gamma^{s-4}
\le \delta_1 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac{c(\delta_1,s,n)}{\rho^{4k}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4k}
\label{EQcy2e2}
\,.
\end{equation}
Combining \eqref{EQcy2e1} with \eqref{EQcy2e2} and then choosing
$\delta_1 = \frac{1}{2c_3}$, where $c_3(n,s)$ is as in the previous Corollary, yields
\begin{equation}
\frac{c_3}{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\le \frac12 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s-4}
+ \frac{\tilde c}{\rho^{4k}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4k}
\label{EQcy2e3}
\, ,
\end{equation}
for some $\tilde c = \tilde c(n,s)$. Using \eqref{EQcy2e3} to estimate the right hand side of Corollary \ref{CY1} we find
\begin{align*}
\frac{d}{dt}E^k_\eta(u) + \frac32E^{k+1}_\eta(u)
&\le
\frac{c_3}{\rho^{8}}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\\&\le
\frac12 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s-4}
+ \frac{\tilde c}{\rho^{4k}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4k}
\,,
\intertext{which, after absorbing the first term on the right into the left, becomes}
\frac{d}{dt}E^k_\eta(u) &+ E^{k+1}_\eta(u)
\le
\frac{c_4}{\rho^{4k}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4k}
\end{align*}
as required.
\end{proof}
\section{Uniqueness}
\label{STuniqueness
In this section we prove that smooth solutions to \eqref{bi} which satisfy
$|\Delta u|^2(\cdot,t) \leq \frac{k_0}{t}$ are uniquely determined by their
initial values.
\begin{theorem}\label{uthm}
Let $v:\mathbb R^n \times [0,T] \to \mathbb R$, $T < \infty$ be a smooth solution to \eqref{bi}
which satisfies
\begin{align}
| \Delta v |^2(x,t) \leq \frac{k_0}{t} \label{introA1u}
\end{align}
for some $k_0 \in \mathbb R$, for all $t \in [0,T]$ and all $x \in \mathbb R^n$ and
\begin{align}
v_0 \equiv 0.
\end{align}
Then $ v \equiv 0$.
\end{theorem}
\begin{proof}
Since $$ \sup_{B_1(0)} \sum_{i=0}^p |\nabla^i v_0|^2 = 0 $$ for any
$p>0$, (c.f. (A2)), Theorem \ref{main1} tells us that $|\Delta v|^2 (0,t) \leq 2N(n,k_0)$ for some
$N = N(n,k_0) \in \mathbb R$ for all $t \leq 1/ N$.
By setting $\tilde v(\cdot,t) = v( \cdot -x_0,t)$ and using Theorem
\ref{main1} for $\tilde v$, we see that $|\Delta v|^2 (x_0,t) \leq N(n,k_0)$
for all $t \leq 1/ N$, for all $x_0 \in \mathbb R^n$.
Corollary \ref{CY2} implies that
\begin{align}
\frac{\partial }{\partial t} E^p_{\eta}(v) + E^{p+1}_{\eta} (v) \leq 2CN
\omega_n\rho^{n-4p} = 2CN
\omega_n\rho^{-3n-4}
\end{align}
where $p$ is now fixed and chosen to be $p(n) = n + 1$, and $\eta $ is
a non-negative cutoff function with $\eta = 1$ on $B_{\rho}(0)$, and
$C = C(n)$. To see this
repeat the argument from the inequality \eqref{step1} up to \eqref{remember}
but use this $v$ (instead of the $v$ appearing there) and use the fact that
$k_0 = 0$, and $|\Delta v|^2 \leq N$ for all $ t\leq 1/N$ for this $v$.
This implies that $E^p_{\eta}(v)(t) \leq 2CN\omega_n\rho^{-3n-4}$ for all $t
\leq 1/N \leq 1$, since
$E^p_{\eta}(v) $ is non-negative, and $E^p_{\eta}(v)( 0) = 0$.
Letting $\rho \to \infty$, we see that $\int_{\mathbb R^n} |\Delta ^p v|^2 =0$ for all $t
\leq 1/N$.
Similarly $\int_{\mathbb R^n} |\Delta ^{p-1} v|^2 = 0$ for all $t
\leq 1/N$.
Now use
\begin{align*}
\frac{\partial }{\partial t} \int_{B_1(0)} |\Delta^{p-2} v|^2 & =
2 \int_{B_1(0)} \Delta^{p} v \Delta^{p-2} v \cr
& \leq \int_{B_1(0)} |\Delta^{p} v|^2 + |\Delta^{p-2} v|^2
= \int_{B_1(0)} |\Delta^{p-2} v|^2
\end{align*}
which tells us, after integrating, that $ \int_{B_1(0)}
|\Delta^{p-2} v|^2 = 0 $ for all $t \leq 1/N$. Differentiating
$\int_{B_1(0)} |\Delta^{p-3} v|^2$ w.r.t. time
and using $\int_{B_1(0)}| \Delta^{p-1} v|^2 = 0$ we obtain, using the same
argument, that $ \int_{B_1(0)}
|\Delta^{p-3} v|^2 = 0 $ for all $t \leq 1/N$. Continuing in this way,
we find that $\int_{B_1(0)}
|\Delta^{l} v|^2 = 0$ for all $0 \leq l \leq p$, for all $t \leq 1/N$.
Similarly, we obtain $\int_{B_1(x_0)}
|\Delta^{l} v|^2 = 0$ for all $0\leq l \leq p$, for all $t \leq 1/N$ for
all $x_0 \in\mathbb R^n$. In
particular, by choosing $l =0$, we see that $ v(\cdot,t) =0$ for all $t \leq
1/N$, $t \leq T$.
Repeating this argument for the function $\tilde v(\cdot,\tilde t) =
v(\cdot,\tilde t + 1/N)$, we see that $v(\cdot,t) =0$ for all $t \leq T$, as required.
\end{proof}
\begin{corollary}\label{uniqueness}
Let $u, v:\mathbb R^n \times [0,T] \to \mathbb R$, $T < \infty$ be smooth solutions to \eqref{bi}
which satisfy
\begin{align}
|\Delta v|^2(x,t) + |\Delta u|^2(x,t) \leq \frac{k_0}{ t} \label{uniqueA1}
\end{align}
for all $t \in [0,T]$ and all $x \in \mathbb R^n$ and
\begin{align}
u_0(\cdot) = v_0(\cdot).
\end{align}
Then $u \equiv v$.
\end{corollary}
\begin{proof}
Set $s = u-v$. Then $s$ satisfies the conditions of Theorem \ref{uthm}. Hence
$s \equiv 0$ as required.
\end{proof}
\section{A Tychonoff-type solution and Non-uniqueness}\label{Tych}
In this section we describe a simple modification to the classical Tychonoff
counterexample, see \cite{Tych}, which establishes non-uniquness for complete solutions of the
polyharmonic heat equation.
We follow the construction given in \cite[Chapter 7, Section 1 (a), pp 211--213]{FritzJohn}.
Let $k\in\mathbb N$ and consider a solution $u:\mathbb R^n\times[0,T]\rightarrow\mathbb R$ to
\begin{align}
(\partial_t-\Delta^k)u &= 0\quad\qquad\text{on}\quad\mathbb R^n\times[0,T]\,,
\label{EQtik1}
\\
u(\cdot,0) &= u_0(\cdot)\,\ \quad\text{on}\quad\mathbb R^n\,.
\label{EQtik2}
\end{align}
We shall construct infinitely many solutions to
\eqref{EQtik1}-\eqref{EQtik2} which have zero as their initial
data.
For functions $g_j:[0,T]\rightarrow\mathbb R$ to be chosen, set
\[
u(x,t) = \sum_{j=0}^\infty g_j(t)x_1^{2jk}\,.
\]
The convergence of this series will be guaranteed by our choice of $g_j$, and verified later.
Differentiating formally, we find
\begin{align*}
\sum_{j=0}^\infty (\partial_tg_j)(t)x_1^{2jk}
&= \partial_tu(x,t)
= \Delta^ku(x,t)
\\
&= \sum_{j=1}^\infty (2jk)(2jk-1)\cdots(2jk-2k+1)g_j(t)x_1^{2jk-2k}
\\
&= \sum_{j=0}^\infty (2jk+2k)(2jk+2k-1)\cdots(2jk+1)g_{j+1}(t)x_1^{2jk}\,
\end{align*}
for all $j \in \mathbb N_0$.
We are thus led to the recurrence relation
\begin{equation}
\label{EQtik3}
(\partial_tg_j) = (2jk+2k)(2jk+2k-1)\cdots(2jk+1)g_{j+1}\,
\end{equation}
for all $j \in \mathbb N_0$.
We set $g_j(t) = \lambda(j,k)g_0^{(j)}(t)$, where $(g_0)^j$ refers to
$j$ temporal derivatives of $g_0$, and $\lambda(j,k)$ is a constant to be determined depending only on
$j,k$. Using this choice of $g_j$, we see that \eqref{EQtik3} is
satisfied, provided that
\begin{equation*}
g_0^{(j+1)}\lambda(j,k) = (2jk+2k)(2jk+2k-1)\cdots(2jk+1)\lambda(j+1,k)g_0^{(j+1)}\,,
\end{equation*}
which for $g_0^{(j+1)}\ne0$ is equivalent to
\begin{equation*}
\frac{\lambda(j,k)}{\lambda(j+1,k)} = (2jk+2k)(2jk+2k-1)\cdots(2jk+1)\,.
\end{equation*}
Using $\lambda(0,k) = 1$, we see that this implies that $\lambda(j,k)
= \frac{1}{(2jk)!}$ for all $j \in \mathbb N_0$ (we use $0!:= 1$) .
Let us now set
\begin{equation*}
g_0(t) = \exp(-t^{-p})
\end{equation*}
for $t>0$ and $p>1$.
\begin{lemma}
\label{LMtikhonovestimate}
There is an absolute constant $\epsilon_0 >0$ and a $p >1$ such that the estimate
\[
\Big|g_0^{(j)}(t)\Big| \le \frac{j!2^j}{t^j}\exp\big(-\epsilon_0(2t)^{-p}\big)
\]
holds for all $t>0$.
\end{lemma}
\begin{proof
Consider the function $h(z)=\exp(-z^{-p})$ for $p>1$.
Since $z^p := \exp(p\rm{Log}(z))$ is analytic on $\mathbb C\backslash(-\infty,0]$,
$h$ is analytic on $\mathbb C\backslash(-\infty,0]$: for polar coordinates $z
= re^{i\theta}$ with $\theta \in (-\pi, \pi)$ we are using $\rm{Log}(z):=
\log(r) + i \theta$, and hence $z^p = r^pe^{ip\theta}$.
For $0<\rho<t$, Cauchy's integral formula on $S_\rho(t+0i) = S_\rho(t)$, the circle in $\mathbb C$ with radius $\rho$
centred at $t+0i$, gives
\[
g_0^{(j)}(t)
= h^{(j)}(t+0i)
= \frac{j!}{2\pi i} \int_{S_\rho(t)} \frac{h(z)}{(z-t)^{j+1}}dz\,.
\]
This gives the estimate
\begin{equation}
\label{EQtik4}
|g_0^{(j)}(t)|
\le \frac{j!}{\rho^j}\sup_{z\in S_\rho(t)} |h(z)|\,.
\end{equation}
In polar coordinates $z = r\exp(i\theta)$, $\theta \in (-\pi,\pi)$ we have
\[
h(z) = \exp\big(-r^{-p}e^{-ip\theta}\big)
= \exp\big(-r^{-p}(\cos(p\theta) - i\sin(p\theta))\big)\,.
\]
Therefore
\begin{equation}
\label{EQrev0}
|h(z)| \le \exp(-r^{-p}\cos p\theta)\,.
\end{equation}
For $z\in S_\rho(t)$, we have
\[
-\frac{\pi}{2} < -\theta_0
\le \theta
\le \tan^{-1}( \rho/\sqrt{t^2-\rho^2} )
=: \theta_0 < \frac{\pi}{2}
\]
Note that $\theta_0$ doesn't depend on $p$.
So we may choose $p>1$ such that $p\theta_0 < \frac{\pi}{2}$: this is
possible since $\theta_0 < \frac{\pi}{2}$.
We then have $\cos (p\theta) \ge \cos(p\theta_0) =: \epsilon_0 > 0$ for all
$\theta \in (-\theta_0, \theta_0)$.
Since $r\le 2t$ we may estimate
\[
-r^{-p} \cos p\theta \le -\epsilon_0 (2t)^{-p}
\]
for all $\theta \in (-\theta_0, \theta_0)$,
which combined with our earlier estimate \eqref{EQrev0} yields
\[
\sup_{z\in S_\rho(t)} |h(z)|
\le \exp\big(-\epsilon_0(2t)^{-p}\big)\,.
\]
Inserting this into the estimate \eqref{EQtik4} and choosing $\rho = t/2$ finishes the proof.
\end{proof
Lemma \ref{LMtikhonovestimate} implies
\begin{align*}
|u(x,t)|
&\leq \sum_{j=0}^\infty |g_j(t)||x_1|^{2jk}
\\
&= \sum_{j=0}^\infty \frac{|g_0^{(j)}(t)|}{(2jk)!} |x_1|^{2jk}
\\
&\le \sum_{j=0}^\infty \frac{j!2^j}{t^j(2jk)!}|x_1|^{2jk}\exp\big(-\epsilon_0(2t)^{-p})
\\
&\le \exp\big(-\epsilon_0(2t)^{-p}\big)
\sum_{j=0}^\infty \frac{j!}{(2jk)!}\Big(\frac{|x_1|^{2k}}{t/2}\Big)^{j}
\\
&\le \exp\big(-\epsilon_0(2t)^{-p} \big)
\sum_{j=0}^\infty \frac{1}{j!}\Big(\frac{|x_1|^{2k}}{t/2}\Big)^{j}
\\
&= \exp\Big(-\frac{\epsilon_0}{(2t)^p} + \frac{|x_1|^{2k}}{t/2}\Big)\,.
\end{align*}
Here we have used $\frac{j!}{(2j)!} \leq \frac{1}{j!}$ for all $j \in
\mathbb N_0$ which may be
seen using induction.
Therefore $u$ is well-defined for every $t>0$. Moreover, $p>1$ implies that the first term above always
dominates for small $t$ and so $u$ converges uniformly to zero on compact subsets of $\mathbb R^n$ as $t\searrow0$.
More precisely, let $K$ be a compact subset of $\mathbb R^n$ with diameter
$d$ and $0 \in K$. Then $|x| \le d$ and for $x\in K$
\[
\lim_{t\searrow0} |u|(x,t)
\le
\lim_{t\searrow0} \exp\Big(-\frac{\epsilon_0}{(2t)^p} + \frac{d^{2k}}{t/2}\Big)
= 0\,.
\]
A similar argument shows that all derivatives of $u$ exist and converge uniformly to
zero on compact subsets of $\mathbb R^n$ as $t\searrow0$.
We explain this in the following.
Assuming $x = x_1$ satisfies $|x| \leq d$ where $d \geq 1$ and taking $s$ spatial derivatives formally we find
\begin{align*}
\left|(\partial_{x})^su(x,t)\right| & = \left| \sum_{j \geq s/(2k)}
(g_0)^j(t)(x)^{2jk-s} \frac{(2jk)(2jk-1) \ldots (2jk -s +1)} {(2jk)!} \right|
\\
& \leq \sum_{j \geq s/(2k)}
\left|(g_0)^j(t)(x)^{2jk-s} \right| \left| \frac{(2jk)(2jk-1) \ldots (2jk -s +1)} {(2jk)!} \right|
\\
& \leq \sum_{j \geq s/(2k)} |g_0^j|(t)|d|^{2jk-s} \frac{ 1}{(2jk-s)!} \\
& \leq \sum_{j \geq s/(2k)} |g_0^j|(t)(|d|^{2k} )^j \frac{ 1}{(2jk-s)!} \\
& \leq \exp( -\epsilon_0(2t)^{-p}) \sum_{j \geq s/(2k)} (
2d^{2k}/t)^j\frac{j!} {(2jk-s)!} \\
\displaybreak[4]
& \leq \exp( -\epsilon_0(2t)^{-p}) \sum_{ \{ j \ | \ 2kj \geq s \} } (
2d^{2k}/t)^j\frac{(kj)!} {(2jk-s)!} \\
&\leq s! \exp( -\epsilon_0(2t)^{-p}) \sum_{ \{ j \ | \ 2kj \geq s \} } \frac{(
2d^{2k}/t)^j}{kj!}\\
& \leq s! \exp( -\epsilon_0(2t)^{-p}) \sum_{ \{ j \ | \ 2kj \geq s \} } \frac{(
2d^{2k}/t)^j}{j!}\\
&\leq s! \exp( -\epsilon_0(2t)^{-p}) \exp( 2d^{2k}/t),
\end{align*}
which goes to zero as $t \searrow 0$.
Here we used that $ \frac{(r!)^2}{(2r-s)!} \leq s!$ for all $r \geq
s$,$r,s \in \mathbb N_0$, which may be verified using induction on $r$.
Since $s$ time derivatives of $u$ are formally given by $2ks$ spatial
derivatives of $u$, we see that all mixed derivatives (space and time) of $u$
exist for $t>0$ and converge uniformly on (spatial) compact sets $K \subset
\mathbb R^n$ to $0$.
By extending $u$ to be zero for all $t\le0$ we have a solution $u\in
C^\infty( \mathbb R^n \times(-\infty,\infty))$ to \eqref{EQtik1}-\eqref{EQtik2} which is
non-zero for $t>0$ and satisfies $u \equiv 0$ for all $t\le 0$.
\section{An example}\label{STexample}
Let $u_0 :\mathbb R^n \to \mathbb R$ be given by $u_0(x_1,x_2, \ldots,x_n):= 1$ if $x_1
>0$, $u_0(x_1,x_2, \ldots,x_n):= 0$ if $x_1 \leq0$.
Setting $u(x,t):= \int_{\mathbb R^n} u_0(x-y) b(y,t) dy$,
with $b: \mathbb R^n \times (0,\infty) \to \mathbb R$ the bi-harmonic heat
kernel on $\mathbb R^n $, we see that the function $u: \mathbb R^n \times
(0,T) \to \mathbb R$ is smooth and solves
$ \frac{\partial }{\partial t} u(x,t) = -\Delta^2u(x,t)$ for all $ t>0$ for all $x \in \mathbb R^n$,
and that $u(\cdot,t) \to u_0(\cdot)$ uniformly on any compact set $K$
contained in $\mathbb R^n \backslash \{ x \in \mathbb R^n | x_1 = 0\}$.
Furthermore, there exists a $ k_0
>0$ such that
for all $s>0$ there exists a $x_s \in \mathbb R^n$ such that
$|\Delta u|^2( x_s,s) = \frac{k_0}{s}$.
The biharmonic heat kernel $b$ is given by
$$b(x,t) = (2\pi)^{-n/2}
t^{-n/4}\int_{\mathbb R^n}e^{ i\langle w,x\rangle t^{-1/4} -|w|^4} dw.
$$
We verify of all these facts below.
We have (see the Appendix of \cite{KL}, and the papers
\cite{GP}, \cite{GG1},\cite{GG2} for further, related and similar estimates)
\begin{align}
\left|\left(\frac{\partial }{\partial t} \right)^l \left(\nabla^k\right) b(x,t)\right| \leq C(k,l,m) (t^{-p(k,l) + m/4} +
t^{ (m-1)/4}) |x|^{-m} \label{HKest}
\end{align}
for all $l,k \in \mathbb N_0, m \in \mathbb N_0$, for some $p(k,l) \in \mathbb N$ for all $x
\in \mathbb R^n$ for all $t >0 $.
This can be seen as follows.
Let $f: \mathbb R^n \to \mathbb R$ be $f(y):= (2 \pi)^{-n/2} \int_{\mathbb R^n}e^{ -i \langle
w, y\rangle-|w|^4} dw,$
so that $b(x,t) = t^{-n/4}f( -\frac{x}{t^{1/4}})$. Then $f$ is the Fourier transform of the function $l: \mathbb R^n \to \mathbb R$,
$l(w):= e^{-|w|^4}$ which is in $\mathcal S$, the so called Schwartz Space (see \cite{SW} Chapter
I.3 where this set of functions is defined and called the space of {\it testing
functions}). Hence $f$ itself is in $\mathcal S$ (see \cite{SW}, Theorem 3.2
Chapter I.3), in particular $|\nabla_{\alpha} f|(x) \leq \frac{c(|\alpha|,m) }{|x|^m }
$ for any $m \in \mathbb N_0$ and any multi-index $\alpha = (\alpha_1, \ldots,
\alpha_n)$: $\alpha_i \in \mathbb N_0 $ for all $i = 1, \ldots, n$, $| \alpha | := \alpha_1 + \alpha_2 + \ldots \alpha_n$, and
we have used the notation $\nabla_{\alpha} f := \nabla_{\alpha_1} \nabla_{\alpha_2} \ldots
\nabla_{\alpha_n} f.$
Using the represenation $b(x,t) = t^{-n/4}f( -\frac{x}{t^{1/4}})$ and the
fact that $f$ is in $\mathcal S$ we get
\begin{align*}
\left|\left(\frac{\partial }{\partial t} \right)^l \left(\nabla^k\right) b(x,t)\right|
&\leq (t^{-p(k,l)} + t^{-1/4})
\sum_{0 \leq |\alpha| \leq k+l} |\nabla_{\alpha} f|\left(-\frac{x}{t^{1/4}}\right) \cr
& \leq (t^{-p(k,l)} + t^{-1/4}) \frac{c(k,l,m) }{ |x/t^{1/4}|^m} \cr
& = c(k,l,m) \frac{t^{-p(k,l) + m/4} +t^{(m-1)/4 } }{ |x|^m}
\end{align*}
which proves the estimate \eqref{HKest} since $m \in \mathbb N_0$ was arbitrary.
This shows that the function $u(x,t):= \int_{\mathbb R^n} u_0(x-y) b(y,t) dy =\int_{\mathbb R^n} u_0(z) b(x-z,t) dz
$ is well defined for any measurable $L^{\infty}$ function $u_0: \mathbb R^n
\to \mathbb R$ for all $t>0$, and is differentiable in time and space for all
$t>0$ for all $x \in \mathbb R^n$ and the derivative is given by
differentiating under the integral sign (in view of the Lebesgue
dominated convergence Theorem):
\begin{align*}
\left(\frac{\partial }{\partial t} \right)^l \left(\nabla^k\right) u(x,t) =\int_{\mathbb R^n} u_0(z) \left(\left(\frac{\partial }{\partial t} \right)^l \left(\nabla^k\right) b\right)(x-z,t) dz.
\end{align*}
Using the fact that $\frac{\partial }{\partial t} b = -\Delta^2 b$ (see below for an explanation), we get
$u: \mathbb R^n \times (0,\infty) \to \mathbb R$ is smooth and satisfies
{$ \frac{\partial }{\partial t} u = -\Delta^2 u$.
Notice also that $\int_{\mathbb R^n} b(x,t) dx = \int_{\mathbb R^n} t^{-n/4} f(
-\frac{x}{t^{1/4}}) dx = \int_{\mathbb R^n} f(-z) dz = \int_{\mathbb R^n} b(z,1)
dz = 1$ (the last equality is explained below).
Hence, for $x \in B_{\epsilon}(z)$ where $ z = (z_1, \ldots, z_n)$ has $z_1
> 2\epsilon$, we have
\begin{align}
|u(x,t) - 1| & = \bigg|\int_{\mathbb R^n} b(x-y,t)(u_0(y) - 1) dy\bigg|
\cr
& = \bigg| \int_{B_{\epsilon}(x)} b(x-y,t) (u_0(y) - 1) dy
\\&\quad + \int_{\mathbb R^n \backslash B_{\epsilon}(x)} b(x-y,t) (u_0(y) - 1) dy \bigg|\cr
& = 0 + \bigg|\int_{\mathbb R^n \backslash B_{\epsilon}(x)} b(x-y,t) (u_0(y) - 1) dy \bigg|\cr
& \leq \int_{\mathbb R^n \backslash B_{\epsilon}(x)} 2|b(x-y,t)| dy \cr
&\leq \int_{\mathbb R^n \backslash B_{\epsilon}(x) } \frac{c(m,n)t^{m-n/4}}{|x-y|^{4m}} dy \cr
&\leq C(\epsilon,m,n)t^{m-n/4} \cr
&\leq C(\epsilon,m,n) t^2
\end{align}
for $m > 2 + n/4$ for $t\leq 1$ which goes to zero as $t \to 0$.
Similarly $|u(x,t)| \leq C(\epsilon,m,n)t^2$ goes to zero for all $x \in B_{\epsilon}(z)$
where $ z = (z_1, \ldots, z_n)$ has $z_1
< -2\epsilon$. Hence $u(\cdot,t) \to u_0$ uniformly on compact sets $K
\subseteq \mathbb R^n \backslash \{ x \in \mathbb R^n \ | \ x_1 = 0\}$.
The definition of $u_0$ and $u$ guarantees that $u(cx,c^4t) = u(x,t)$
for all $c,t>0$. We verify this now. Notice first that
\begin{align}
b(cx,c^4t) & = (2\pi)^{-n/2}
(c^4t)^{-n/4}\int_{\mathbb R^n}e^{ i(c^4t)^{-1/4} \langle w,cx\rangle-|w|^4} dw\cr
& =c^{-n} (2\pi)^{-n/2}\int_{\mathbb R^n}e^{ i t^{-1/4} \langle
w,x\rangle-|w|^4} dw
\end{align}
that is $b(cx,c^4t) = c^{-n}b(x,t) $ for all $ x \in \mathbb R^n$ for all $c>0$.
Also, the definition of $u_0$ guarantees that $u_0(cz) = u_0(z)$ for
all $z \in \mathbb R^n$ and all $c>0$.
Making a change of variable $y = cw$ in the definition of $u$, and
then using $b(cw,c^4t) = c^{-n}b(w,t) $ and the property of $u_0$ just
mentioned, we calculate
\begin{align}
u(cx,c^4t):= \int_{\mathbb R^n} u_0(cx-y) b(y,c^4t) dy
&= \int_{\mathbb R^n} u_0(cx-cw) b(cw,c^4t) c^n dw \cr
& = \int_{\mathbb R^n} u_0(x-w) b(w,t) dw \cr
& = u(x,t).
\end{align}
There must exist a point $(x_0,t_0) \in \mathbb R^n \times \mathbb R^+$ with
$\Delta u(x_0,t_0) \neq 0$: if not, then $\frac{\partial }{\partial t} u = -\Delta^2 u = 0 $ for
all $t >0$, and hence $u(x,t) = u(x,s)$ for all $0<s<t$, and hence ,
using $u \to u_0$ on $\mathbb R^n \backslash \{ x \in \mathbb R^n | x_1 = 0\}$ as $t
\searrow 0$
as explained above, we have $u(x,t) = 1$ on $\mathbb R^n \cap \{ x \in
\mathbb R^n | x_1 > 0\}$, $u(x,t) = 0$ on $\mathbb R^n \cap \{ x \in \mathbb R^n | x_1 <
0\}$ for all $t>0$, which contradicts the fact that $u(\cdot,t): \mathbb R^n
\to \mathbb R$ is smooth for $t>0$.
So there exists $(x_0,t_0) \in \mathbb R^n \times \mathbb R^+$ with $|\Delta u(x_0,t_0)|^2 \neq
0$. Now $u(cx,c^4,t) = u(x,t)$ for all $t>0$, for all $x \in \mathbb R^n$ implies that
(take the Laplacian w.r.t $x$ of both sides) $c^2 (\Delta u)(cx,c^4t) = (\Delta
u)(x,t)$ which implies $|\Delta u|^2(cx,c^4t) = \frac{1}{c^4} |\Delta u|^2(x,t)$.
In particular, choosing $t = t_0$, $c^4 = (s/t_0)$ and $x = x_0$ we find
\[
|\Delta u|^2((s/t_0)^{1/4} x_0, s) = \frac{t_0}{s} |\Delta
u|^2(x_0,t_0)
\]
and hence
$|\Delta u|^2( x_s,s) = \frac{k_0}{s}$ where $k_0 = t_0 |\Delta
u(x_0,t_0)|^2 \neq 0$ and $x_s = (s/t_0)^{1/4} x_0$.
Using an almost identical argument, we see that for all $t>0$, there
must be points $y(t) \in \mathbb R^n$ such that
$(\Delta^2 u)( y(t),t) = \frac{k_1}{t}$ for some fixed $k_1 \in \mathbb R$, $k_1
\neq0 $.
The fact that $\frac{\partial }{\partial t} b = -\Delta^2 b$ can be seen as follows. Using
Theorem 1.7 of Chapter I.1 in \cite{SW}, we have
$-|x|^4 e^{-|x|^4 t} = \frac{\partial }{\partial t} (e^{-|x|^4 t}) =(\frac{\partial }{\partial t} \widehat b)(x,t) =
\widehat{(\frac{\partial }{\partial t} b)}(x,t) = \widehat{( -\Delta^2 b)}(x,t)$, and hence, taking
the inverse of the Fourier transform, we get $(\frac{\partial }{\partial t} b) = -\Delta^2 b$ (note that $(\frac{\partial }{\partial t} \widehat b)(x,t) =
\widehat{(\frac{\partial }{\partial t} b)}(x,t) $ is true in view of the Lebesgue dominated
convergence Theorem and the estimates \eqref{HKest} and the inverse of
the Fourier transform exists in view of Corollary I.21 in I.1 of
\cite{SW} and the fact that $b$ is in $\mathcal S$).
The fact that $\int_{\mathbb R^n} b(z,1)
dz = 1$
may be seen by looking at how $b$ was derived:
Let $ u_0: \mathbb R^n \to \mathbb R$ be a smooth function which is equal to $1$ on
$B_1(0)$ and has compact support on $B_2(0)$.
Hence $u_0$ is in $\mathcal S$, and the Fourier transform
$\widehat{u_0} $ of $u_0$ is also in $\mathcal S$.
We only take the Fourier transform in the space direction in that
which follows.
Write $u(x,t) = (b(\cdot,t) * u_0)(x)$ so $\frac{\partial }{\partial t} u = -\Delta^2 u$ as
explained above, and
\begin{align}
\widehat u (x,t) & = \widehat b (x,t) \widehat{u_0}(x)
= e^{-t|x|^4} \widehat u_0(x)
\end{align}
(see Theorem 1.4 in I.1 os \cite{SW}) and hence $\widehat u(\cdot,t) \to \widehat{u_0}(\cdot)$
in the $L^2$ sense as $t \searrow 0$.
But this implies $u(\cdot,t) \to u_0(\cdot)$ in the $L^2$ sense as $t \searrow 0$,
in view of the fact that the
$L^2$ norm is preserved for the Fourier transform (and the inverse
Fourier transform) of a function in
$L^2\cap L^1$ (see Theorem 2.1 and Theorem 2.4 in Chapter I.2 of \cite{SW}).
In particular this shows $\int_{\mathbb R^n} b =1$:
If $1 \neq c_0 := \int_{\mathbb R^n} b \neq 0$, then for $x \in B_{1/2}(0)$
we have
\begin{align}
|u(x,t) - 1/c_0| & =
\bigg|\int_{\mathbb R^n} b(x-y,t)(u_0(y) - 1) dy\bigg|
\cr
& = \bigg|\int_{\mathbb R^n \backslash B_1(0)} b(x-y,t)(u_0(y) - 1) dy\bigg|
\cr
& = \bigg|\int_{\mathbb R^n \backslash B_1(0)} b(x-y,t)(u_0(y) - 1) dy\bigg|
\cr
& \leq C \int_{\mathbb R^n \backslash B_1(0)} |b(x-y,t)| dy\cr
& = C \int_{\mathbb R^n \backslash B_1(x)} |b(z,t)| dz\cr
& \leq C \int_{\mathbb R^n \backslash B_{1/2}(0)} |b(z,t)| dz\cr
& \to 0
\end{align}
as $t \searrow 0$
in view of \eqref{HKest},
which shows $u(\cdot,t)$ converges uniformly in the supremum norm to
$(1/c_0) \neq 1$ on $B_{1/2}(0)$ as $t \searrow 0$,
which contradicts the fact that $u(\cdot,t)$ converges to $u_0$ in the
$L^2$ norm as $t \searrow 0$.
Similarly, if $ \int_{\mathbb R^n} b = 0 $, one shows $u(x,t) \to 0$ uniformly in the supremum norm
on $B_{1/2}(0)$ as $t \searrow 0$, which contradicts the fact that $u(\cdot,t)$ converges to $u_0$ in the
$L^2$ norm as $t \searrow 0$.
\section{Appendix
We require a rather specific form of the standard interpolation inequalities (see for example \cite[Theorems
7.25--7.28]{GT}). We have thus provided
proofs and precise statements of that which we need here in the appendix.
\begin{lemma}\label{L2est}
For any smooth function $\varphi:B_1(0) \to \mathbb R$ we have
\begin{equation}
\int_{B_{1-s\sigma}} |\nabla^{2s} \varphi|^2 + |\nabla^{2s-1} \varphi|^2\leq c(s,\sigma) \int_{B_1} |\Delta^s
\varphi|^2 + |\Delta^{s-1}\varphi|^2 + \ldots + |\varphi|^2 \label{inducts}
\end{equation}
\end{lemma}
for any $s \in \mathbb N$ and $1>\sigma>0$, as long as $1-s\sigma >0$.
\begin{proof
We show the inequality \eqref{inducts} for arbitrary smooth $\varphi:B_1(0) \to \mathbb R$ using induction.
Step 1: $L^2$-theory (see for example \cite[Theorem 1, Section 6.3.1]{Evans})
tells us that for an arbitrary smooth function $\varphi:B_1 \to \mathbb R$
\begin{equation}
\int_{B_{1-\sigma}} |\nabla^{2} \varphi|^2 + |\nabla \varphi|^2\leq c(\sigma)
\int_{B_1} |\Delta \varphi|^2 + |\varphi|^2
\end{equation}
as required.
Inductive Step:
Let $\alpha = (\alpha_1, \ldots, \alpha_n)$ be a multi-index with
$0\leq \alpha_i \leq 2s$ and $\sum_{i=1}^n \alpha_i = 2s$.
We use the notation $\nabla_{\alpha} \varphi := \nabla_{\alpha_1} \nabla_{\alpha_2}
\ldots \nabla_{\alpha_n} \varphi.$
Assume that statement \eqref{inducts} is true for some $s \in \mathbb N$.
Again, $L^2$ theory, see for example \cite{Evans} Theorem
1, Section 6.3.1, tells us that
\begin{align*}
\int_{B_{1-(s+1)\sigma}} &|\nabla^{2} \nabla_{\alpha} \varphi|^2 + |\nabla \nabla_{\alpha} \varphi|^2
\\& \leq
a(s,\sigma) \int_{B_{1-s\sigma}} |\Delta (\nabla_{\alpha} \varphi)|^2 + |\nabla_{\alpha}
\varphi|^2\cr
&= a(s,\sigma) \int_{B_{1-s\sigma}} |\nabla_{\alpha} (\Delta \varphi)|^2 + |\nabla_{\alpha}
\varphi|^2\cr
& \leq a(s,\sigma) c(s,\sigma) \int_{B_1} |\Delta^s
(\Delta \varphi) |^2 + |\Delta^{s-1} ( \Delta \varphi)|^2 + \ldots + |\Delta
\varphi|^2\cr
& \ \ \ + a(s,\sigma) c(s,\sigma) \int_{B_1} |\Delta^s
\varphi|^2 + |\Delta^{s-1}\varphi|^2 + \ldots + |\varphi|^2 , \cr
& = \tilde a (s,\sigma) \int_{B_1} |\Delta^{s+1}
\varphi|^2 + |\Delta^{s}\varphi|^2 + \ldots + |\varphi|^2
\end{align*}
where in the second last line (the inequality) we have used the induction hypothesis
applied to the functions $\Delta \varphi$ and $\varphi$.
By summing up over all possible $\alpha$ (the number of possible $\alpha$
is a constant depending on $n$ and $s$) we see that
\begin{align*}
\int_{B_{1-(s+1)\sigma}} |\nabla^{2s +2} \varphi|^2 + |\nabla^{2s +1} \varphi|^2
& \leq c (s +1,\sigma) \int_{B_1} |\Delta^{s+1}
\varphi|^2 + |\Delta^{s}\varphi|^2 + \ldots + |\varphi|^2
\end{align*}
as required.
This completes the proof by induction.
\end{proof
\begin{lemma}
Suppose $u\in C^\infty_{\text{loc}}(\mathbb R^n)$, $k\in\mathbb N$, $s>8$, and $\gamma$, $c_\gamma$ are as in \eqref{EQgamma}.
For any $\delta_0>0$ we have
\label{LMinterp1}
\begin{equation*}
\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\le \delta_0\rho^4 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^s
+ \frac{c_{\delta_0}}{\rho^4}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,.
\end{equation*}
where $c_{\delta_0}$ is an absolute constant given by
\[
c_{\delta_0}
= c_{\delta_0}(\delta_0,s,n)
= 2^4(\delta_0^{-1} + 2^9s^4c_{\gamma}^4)\,.
\]
\end{lemma}
\begin{proof
Integrating by parts,
\begin{align*}
\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
&= -\int_{\mathbb R^n} (\nabla_i\Delta^ku)(\nabla_i\Delta^{k-1}u)\gamma^{s-4}
\\&\qquad
-(s-4)\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k-1}u)(\nabla_i\gamma)\gamma^{s-5}
\\
&= \int_{\mathbb R^n} (\Delta^{k+1}u)(\Delta^{k-1}u)\gamma^{s-4}
\\*&\qquad
+ (s-4)\int_{\mathbb R^n} (\nabla_i\Delta^ku)(\Delta^{k-1}u)(\nabla_i\gamma)\gamma^{s-5}
\\*&\qquad
-(s-4)\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k-1}u)(\nabla_i\gamma)\gamma^{s-5}
\\
&= \int_{\mathbb R^n} (\Delta^{k+1}u)(\Delta^{k-1}u)\gamma^{s-4}
\\*&\qquad
-2(s-4)\int_{\mathbb R^n} (\Delta^ku)(\nabla_i\Delta^{k-1}u)(\nabla_i\gamma)\gamma^{s-5}
\\*&\qquad
- (s-4)\int_{\mathbb R^n} (\Delta^ku)(\Delta^{k-1}u)(\Delta\gamma)\gamma^{s-5}
\\*&\qquad
- (s-4)(s-5)\int_{\mathbb R^n} (\Delta^ku)(\Delta^{k-1}u)|\nabla\gamma|^2\gamma^{s-6}\,.
\end{align*}
Estimating the right hand side with Young's inequality (estimate \eqref{EQcauchy}), we have for any $\delta_i>0$
\begin{align*}
\int_{\mathbb R^n}& |\Delta^ku|^2\gamma^{s-4}
\\
&\le
\delta_1\rho^4\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac{1}{4\delta_1\rho^4}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\\*&\qquad
+ \delta_2\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
+ \frac{c_\gamma^2(s-4)^2}{\delta_2\rho^2}\int_{\mathbb R^n} |\nabla\Delta^{k-1}u|^2\gamma^{s-6}
\\*&\qquad
+ \delta_3\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
+ \frac{c_\gamma^2(s-4)^2}{4\delta_3\rho^4}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-6}
\\*&\qquad
+ \delta_4\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
+ \frac{c_\gamma^4(s-4)^2(s-5)^2}{4\delta_4\rho^4}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,,
\end{align*}
which upon absorption gives
\begin{align}
\int_{\mathbb R^n}& |\Delta^ku|^2\gamma^{s-4}
\notag\\
&\le
4\delta_1\rho^4\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac{16c_\gamma^2(s-4)^2}{\rho^2}\int_{\mathbb R^n} |\nabla\Delta^{k-1}u|^2\gamma^{s-6}
\notag\\*&\qquad
+ \frac{1}{\rho^4}\bigg(
\frac{1}{\delta_1}
+ 4c_\gamma^2(s-4)^2
+ 4c_\gamma^4(s-4)^2(s-5)^2
\bigg)\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,,
\label{EQlminterpeq1}
\end{align}
where we have chosen $\delta_2 = \delta_3 = \delta_4 = \frac{1}4$, and
we used $\gamma^{s-6}(\cdot) \leq \gamma^{s-8}(\cdot)$, which follows in
view of $0 \leq \gamma \leq 1$ and $s-6>s-8\geq 1$.
For the second term we integrate by parts and estimate using Young's inequality to obtain
\begin{align*}
\int_{\mathbb R^n} &|\nabla\Delta^{k-1}u|^2\gamma^{s-6}
\\
&=
- \int_{\mathbb R^n} (\Delta^{k}u)(\Delta^{k-1}u)\gamma^{s-6}
- (s-6)\int_{\mathbb R^n} (\Delta^{k-1}u)(\nabla_i\Delta^{k-1}u)(\nabla_i\gamma)\gamma^{s-7}
\\
&\le \frac{\rho^2}{64c_{\gamma}^2(s-4)^2}\int_{\mathbb R^n} |\Delta^{k}u|^2\gamma^{s-4}
+ \frac{16c_{\gamma}^2(s-4)^2}{\rho^2}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}
\\*&\qquad
+ \frac{1}{2}\int_{\mathbb R^n} |\nabla\Delta^{k-1}u|^2\gamma^{s-6}
+ \frac{c_\gamma^2(s-6)^2}{2\rho^2}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,.
\end{align*}
Absorbing the third term on the right into the left yields
\begin{align*}
&\int_{\mathbb R^n} |\nabla\Delta^{k-1}u|^2\gamma^{s-6}
\le \frac{\rho^2}{32c_{\gamma}^2(s-4)^2}\int_{\mathbb R^n} |\Delta^{k}u|^2\gamma^{s-4}
\\*&\qquad
+ \frac{c_{\gamma}^2}{\rho^2}
\bigg(32(s-4)^2 + (s-6)^2\bigg)
\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,,
\intertext{and so}
\frac{16c_{\gamma}^2(s-4)^2}{\rho^2}&\int_{\mathbb R^n} |\nabla\Delta^{k-1}u|^2\gamma^{s-6}
\le \frac12\int_{\mathbb R^n} |\Delta^{k}u|^2\gamma^{s-4}
\\*&\qquad
+ \frac{16c_{\gamma}^4(s-4)^2}{\rho^4}\bigg(32(s-4)^2
+ (s-6)^2\bigg)
\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,.
\end{align*}
Combining the above with \eqref{EQlminterpeq1} we have
\begin{align*}
\int_{\mathbb R^n}& |\Delta^ku|^2\gamma^{s-4}
\le
4\delta_1\rho^4\int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^{s}
+ \frac12\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\notag\\*&\qquad
+ \frac{1}{\rho^4}\bigg(
{16c_{\gamma}^4(s-4)^2}\bigg(32(s-4)^2 + (s-6)^2\bigg)
\notag\\*&\qquad\qquad
+ \frac{1}{\delta_1}
+ 4c_\gamma^2(s-4)^2
+ 4c_\gamma^4(s-4)^2(s-5)^2
\bigg)\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,,
\end{align*}
Absorbing the second term on the right into the left, multiplying through by 2 and choosing $\delta_1 = \frac{\delta_0}{8}$ yields the estimate
\begin{equation*}
\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\le \delta_0\rho^4 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^s
+ \frac{\tilde{c}_{\delta_0}}{\rho^4}\int_{\mathbb R^n} |\Delta^{k-1}u|^2\gamma^{s-8}\,
\end{equation*}
where
\begin{align*}
\tilde{c}_{\delta_0}
&=
32c_{\gamma}^4(s-4)^2\left(32(s-4)^2 + (s-6)^2\right)
+ \frac{16}{\delta_0} + 8(s-4)^2\left(c_{\gamma}^2 + c_{\gamma}^4(s-5)^2\right)
\\
&= \frac{16}{\delta_0} + 8(s-4)^2c_{\gamma}^2\left(1 + c_{\gamma}^2(s-5)^2 + 4c_{\gamma}^2\left(32(s-4)^2 + (s-6)^2\right)\right)
\,.
\end{align*}
Since $s>8$ and $c_{\gamma}\ge1$ we have
\begin{align*}
\tilde{c}_{\delta_0}
&\le \frac{16}{\delta_0} + 8s^2c_{\gamma}^2\left(1 + c_{\gamma}^2s^2 + 4c_{\gamma}^2\left(32s^2 + s^2\right)\right)
\\
&\le \frac{2^4}{\delta_0} + 2^3s^4c_{\gamma}^2\left(1 + c_{\gamma}^2 + 2^2c_{\gamma}^2\left(2^5 + 1\right)\right)
\\
&\le 2^4(\delta_0^{-1} + 2^9s^4c_{\gamma}^4)
\\
&:= c_{\delta_0}\,,
\end{align*}
yielding the constant claimed.
\end{proof
\begin{corollary}
\label{LMinterp2}
Suppose $u\in C^\infty_{\text{loc}}(\mathbb R^n)$, $k\in\mathbb N$, and $\gamma$, $c_\gamma$ are as in \eqref{EQgamma}.
For all $k\in\mathbb N$, $s > 4k$,
\begin{equation*}
\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\le \delta\rho^4 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^s
+ \frac{c}{\rho^{4k-4}}\int_{\mathbb R^n} |\Delta u|^2\gamma^{s-4k}\,,
\end{equation*}
where $c(\delta,s,n)< \infty$ is a constant depending on $\delta,s,n$.
\end{corollary}
\begin{proof}
We shall proceed by induction in $k \in \mathbb N$. We wish to show that
\begin{equation}
E^k_{\gamma^{s-4}}(u)
\le {\hat \delta}\rho^4 E^{k+1}_{\gamma^s}(u)
+ \frac{\hat c_{\hat \delta,k,s} }{\rho^{4k-4}}E^{1}_{\gamma^{s-4k}}(u)
\,.
\label{EQindd2}
\end{equation}
is true for all $s > 4k$ for all $\hat \delta > 0$, where
$\hat c_{\hat \delta,k,s}= \hat c(\hat \delta,s,k,n)$.
Let $k=1$.
Then \eqref{EQindd2} is true for all $s > 4k$ for all $\delta>0$ trivially with the
choice of $ \hat c_{\hat \delta,k,s}= 1$.
Assume \eqref{EQindd2} for some fixed $k \in \mathbb N$, and let $s >
4(k+1)$,$\delta >0$ be arbitrary. Then $s > 8$ and
Lemma \ref{LMinterp1} gives the estimate
\begin{equation}
E^l_{\gamma^{s-4}}(u)
\le
\delta\rho^4 E^{l+1}_{\gamma^s}(u)
+ \frac{c_{\delta,l,s}}{\rho^4} E^{l-1}_{\gamma^{s-8}}(u)
\label{EQindd1}
\end{equation}
for any $l\in\mathbb N$, where $c_{\delta,l,s} = c(\delta,l,s,n)$.
Using this inequality with $l = k+1 \in \mathbb N$ and then \eqref{EQindd2} we have
\begin{align*}
E^{k+1}_{\gamma^{s-4}}(u)
&\le
\delta\rho^4 E^{k+2}_{\gamma^s}(u)
+ \frac{c_{\delta, {k+1},s } } {\rho^4} E^{k}_{\gamma^{s-8}}(u)
\\
&\le
\delta\rho^4 E^{k+2}_{\gamma^s}(u)
+ \hat{\delta} c_{\delta, k+1,s} E^{k+1}_{\gamma^{s-4}}(u)
+ \frac{\hat c_{\hat \delta, k,s-4} c_{\delta,k+1,s}}{\rho^{4k}}E^{1}_{\gamma^{s-4-4k}}(u)
\,.
\end{align*}
where here we used the fact that $s > 4(k+1) = 4k +4$ implies that
$\tilde s = s -4 > 4k$ and so \eqref{EQindd2} is valid with $s$
replaced by $\tilde s$.
Choosing
$\hat{\delta} c_{\delta,{k+1},s} = \frac12$ and absorbing we obtain
\begin{align*}
E^{k+1}_{\gamma^{s-4}}(u)
&\le
2\delta \rho^4 E^{k+2}_{\gamma^s}(u)
+ 2\frac{\hat c_{\hat \delta, k,s-4} c_{\delta,k+1,s} }{ \rho^{4k} }E^{1}_{\gamma^{s-4-4k}}(u)
\,.
\end{align*}
which gives us the result, as $\delta>0$ was arbitrary.
Note that we can ensure the constant only depends on $n,s$ and not $k$
by taking the supremum of the constants we obtained in this argument,
where this supremum is taken over all $4k<s$, $k \in \mathbb N$, for a fixed $s \in
\mathbb N$.
\end{proof}
\begin{comment
\begin{corollary}
Suppose $u\in C^\infty_{\text{loc}}(\mathbb R^n)$, $k\in\mathbb N$, $s>8$, and $\gamma$, $c_\gamma$ are as in \eqref{EQgamma}.
For all $k\in\mathbb N$, $q\in\mathbb N\cap[1,k]$ and $s\ge4+4q$,
\label{LMinterp2}
\begin{equation*}
\int_{\mathbb R^n} |\Delta^ku|^2\gamma^{s-4}
\le \delta\rho^4 \int_{\mathbb R^n} |\Delta^{k+1}u|^2\gamma^s
+ \frac{c}{\rho^{4q}}\int_{\mathbb R^n} |\Delta^{k-q}u|^2\gamma^{s-4-4q}\,,
\end{equation*}
where $c$ is an absolute constant depending only on $s$, $c_{\gamma}$ and $\delta$.
\end{corollary}
\begin{proof}
Lemma \ref{LMinterp1} gives the estimate
\begin{equation}
E^k_{\gamma^{s-4}}(u)
\le
\delta_{0,k}\rho^4 E^{k+1}_{\gamma^s}(u)
+ \frac{c_{\delta_0,k}}{\rho^4} E^{k-1}_{\gamma^{s-8}}(u)
\label{EQind1}
\end{equation}
for any $k\in\mathbb N$.
The constant does \emph{not} depend on $k$, but, since we wish to apply Lemma \ref{LMinterp1} several times over with
different $\delta_0$, we add the subscript $k$ to $\delta_0$ to denote which application we are referring to.
Suppose for some $q\in[1,k]\cap\mathbb N$ we have
\begin{equation}
E^k_{\gamma^{s-4}}(u)
\le
\delta_{0,k}\rho^4 E^{k+1}_{\gamma^s}(u)
+ \frac{c_{\delta_0,k}}{\rho^{4q}} E^{k-q}_{\gamma^{s-4-4q}}(u)
\,.
\label{EQind2}
\end{equation}
The estimate \eqref{EQind1} implies the following for any $j \ge q_0-k+1$:
\begin{equation}
E^{k-q_0+j}_{\gamma^{s-4}}(u)
\le
\delta_{0,(k-q_0+j)}\rho^4 E^{k-q_0+j+1}_{\gamma^s}(u)
+ \frac{c_{\delta_0,(k-q_0+j)}}{\rho^4} E^{k-q_0+j-1}_{\gamma^{s-8}}(u)
\label{EQind4}
\,.
\end{equation}
We claim that the following holds
\begin{equation}
E^{k-q_0}_{\gamma^{s-4}}(u)
\le
\delta_{0,(k-q_0+j)}\rho^{4+4j} E^{k-q_0+j+1}_{\gamma^{s+4j}}(u)
+ \frac{c_{\delta_0,(k-q_0)}}{\rho^4} E^{k-q_0-1}_{\gamma^{s-8}}(u)
\,.
\label{EQind5}
\end{equation}
Let us quickly prove \eqref{EQind5} by induction.
Suppose \eqref{EQind5} holds for $j$; then applying \eqref{EQind4} in combination with the inductive
hypothesis \eqref{EQind5} yields
\begin{align}
E^{k-q_0}_{\gamma^{s-4}}(u)
&\le
\delta_{0,(k-q_0+j)}\rho^{4+4j} E^{k-q_0+j+1}_{\gamma^{s+4j}}(u)
+ \frac{c_{\delta_0,(k-q_0)}}{\rho^4} E^{k-q_0-1}_{\gamma^{s-8}}(u)
\notag
\\
&\le
\delta_{0,(k-q_0+j)}\rho^{4+4j}
\bigg(
\delta_{0,(k-q_0+j+1)}\rho^4E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
\notag
\\&\qquad\qquad\qquad\qquad\qquad
+ \frac{c_{\delta_0,(k-q_0+j+1)}}{\rho^4} E^{k-q_0+j}_{\gamma^{s+4j-4}}(u)
\bigg)
\notag
\\&\qquad
+ \frac{c_{\delta_0,(k-q_0)}}{\rho^4} E^{k-q_0-1}_{\gamma^{s-8}}(u)
\notag
\\
&=
\delta_{0,(k-q_0+j)}
\delta_{0,(k-q_0+j+1)}\rho^{8+4j}E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
\notag\\&\qquad
\label{EQind6}
+ \delta_{0,(k-q_0+j)}
{c_{\delta_0,(k-q_0+j+1)}}\rho^{4j} E^{k-q_0+j}_{\gamma^{s+4j-4}}(u)
\\&\qquad
+ \frac{c_{\delta_0,(k-q_0)}}{\rho^4} E^{k-q_0-1}_{\gamma^{s-8}}(u)
\notag
\end{align}
Using \eqref{EQind2} and then \eqref{EQind4} we estimate
\begin{align}
E^{k-q_0+j}_{\gamma^{s+4j-4}}(u)
&\le
\delta_{0,(k-q_0+j)}\rho^4 E^{k-q_0+j+1}_{\gamma^{s+4j}}(u)
+ \frac{c_{\delta_0,k-q_0+j}}{\rho^{4j}} E^{k-q_0}_{\gamma^{s-4}}(u)
\notag\\&\le
\delta_{0,(k-q_0+j)}\rho^4
\bigg(
\delta_{0,(k-q_0+j+1)}\rho^4E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
\notag
\\&\qquad\qquad\qquad\qquad\qquad
+ \frac{c_{\delta_0,(k-q_0+j+1)}}{\rho^4} E^{k-q_0+j}_{\gamma^{s+4j-4}}(u)
\bigg)
\notag\\&\qquad
+ \frac{c_{\delta_0,k-q_0+j}}{\rho^{4j}} E^{k-q_0}_{\gamma^{s-4}}(u)
\notag\\&=
\delta_{0,(k-q_0+j)}
\delta_{0,(k-q_0+j+1)}\rho^8E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
\notag
\\&\qquad
+ \delta_{0,(k-q_0+j)}
{c_{\delta_0,(k-q_0+j+1)}} E^{k-q_0+j}_{\gamma^{s+4j-4}}(u)
\notag\\&\qquad
+ \frac{c_{\delta_0,k-q_0+j}}{\rho^{4j}} E^{k-q_0}_{\gamma^{s-4}}(u)
\,.
\label{EQind7}
\end{align}
For all $\delta_{0,(k-q_0+j)}$ such that $\delta_{0,(k-q_0+j)}{c_{\delta_0,(k-q_0+j+1)}} \le \frac12$ we may absorb the
second term on the right into the left hand side in order to obtain
\begin{align}
E^{k-q_0+j}_{\gamma^{s+4j-4}}(u)
\le
\frac{\delta_{0,(k-q_0+j+1)}}{c_{\delta_0,(k-q_0+j+1)}}
\rho^8E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
+ 2\frac{c_{\delta_0,k-q_0+j}}{\rho^{4j}} E^{k-q_0}_{\gamma^{s-4}}(u)
\,.
\label{EQind8}
\end{align}
Combining \eqref{EQind8} with \eqref{EQind6} proves the estimate
\begin{align}
&E^{k-q_0}_{\gamma^{s-4}}(u)
\\
&\le
\delta_{0,(k-q_0+j)}
\delta_{0,(k-q_0+j+1)}\rho^{8+4j}E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
\notag\\&\quad
+ \delta_{0,(k-q_0+j)}
{c_{\delta_0,(k-q_0+j+1)}}\rho^{4j}
\bigg(
\frac{\delta_{0,(k-q_0+j+1)}}{c_{\delta_0,(k-q_0+j+1)}}
\rho^8E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
\notag\\&\hskip+5cm
+ 2\frac{c_{\delta_0,k-q_0+j}}{\rho^{4j}} E^{k-q_0}_{\gamma^{s-4}}(u)
\bigg)
\notag\\&\qquad
+ \frac{c_{\delta_0,(k-q_0)}}{\rho^4} E^{k-q_0-1}_{\gamma^{s-8}}(u)
\notag\\&=
2\delta_{0,(k-q_0+j)}
\delta_{0,(k-q_0+j+1)}\rho^{8+4j}E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
\notag\\&\quad
+ 2\delta_{0,(k-q_0+j)}
{c_{\delta_0,(k-q_0+j+1)}}
{c_{\delta_0,(k-q_0+j)}} E^{k-q_0}_{\gamma^{s-4}}(u)
\notag\\&\qquad
+ \frac{c_{\delta_0,(k-q_0)}}{\rho^4} E^{k-q_0-1}_{\gamma^{s-8}}(u)
\notag
\end{align}
For all $\delta_{0,(k-q_0+j)}$ such that $2\delta_{0,(k-q_0+j)}{c_{\delta_0,(k-q_0+j+1)}}{c_{\delta_0,(k-q_0+j)}} \le
\frac12$ we may absorb the second term on the right into the left hand side in order to obtain
\begin{align}
E^{k-q_0}_{\gamma^{s-4}}(u)
\le
4\delta_{0,(k-q_0+j)}
\delta_{0,(k-q_0+j+1)}\rho^{8+4j}E^{k-q_0+j+2}_{\gamma^{s+4j+4}}(u)
+ 2\frac{c_{\delta_0,(k-q_0)}}{\rho^4} E^{k-q_0-1}_{\gamma^{s-8}}(u)
\,.
\label{EQind9}
\end{align}
By choosing now $\delta_{0,(k-q_0+j+1)}$ appropriately we conclude our proof of \eqref{EQind5}.
Let us prove \eqref{EQind2} by induction on $k$.
We wish to show for any $\delta>0$ there exists a constant $c = c(\delta,s,c_{\gamma})$ such that
\begin{equation}
E^{k+1}_{\gamma^{s-4}}(u)
\le
\delta\rho^4 E^{k+2}_{\gamma^s}(u)
+ \frac{c}{\rho^{4q}} E^{k+1-q}_{\gamma^{s-4-4q}}(u)
\,.
\label{EQind12}
\end{equation}
First, \eqref{EQind1} gives
\begin{equation}
E^{k+1}_{\gamma^{s-4}}(u)
\le
\delta_{0,(k+1)}\rho^4 E^{k+2}_{\gamma^s}(u)
+ \frac{c_{\delta_0,(k+1)}}{\rho^4} E^{k}_{\gamma^{s-8}}(u)
\,.
\label{EQind10}
\end{equation}
Combining this with the inductive hypothesis \eqref{EQind2} gives
\begin{align*}
E^{k+1}_{\gamma^{s-4}}(u)
&\le
\delta_{0,(k+1)}\rho^4 E^{k+2}_{\gamma^s}(u)
+ \frac{c_{\delta_0,(k+1)}}{\rho^4}
\bigg(
\delta_{0,(k)}\rho^4 E^{k+1}_{\gamma^{s-4}}(u)
+ \frac{c_{\delta_0,(k)}}{\rho^{4q}} E^{k-q}_{\gamma^{s-8-4q}}(u)
\bigg)
\\
&=
\delta_{0,(k)}\delta_{0,(k+1)}\rho^4 E^{k+2}_{\gamma^s}(u)
+ {\delta_{0,(k)}c_{\delta_0,(k+1)}}
E^{k+1}_{\gamma^{s-4}}(u)
+ \frac{{c_{\delta_0,(k+1)}}c_{\delta_0,(k)}}{\rho^{4q+4}} E^{k-q}_{\gamma^{s-8-4q}}(u)
\,.
\end{align*}
Choosing ${\delta_{0,(k)}c_{\delta_0,(k+1)}} = \frac12$ and then absorbing the second term on the right hand side into the
left hand side gives
\begin{align*}
E^{k+1}_{\gamma^{s-4}}(u)
&\le
\frac{\delta_{0,(k+1)}}{c_{\delta_0,(k+1)}}\rho^4 E^{k+2}_{\gamma^s}(u)
+ 2\frac{{c_{\delta_0,(k+1)}}c_{\delta_0,(k)}}{\rho^{4q+4}} E^{k-q}_{\gamma^{s-8-4q}}(u)
\,.
\end{align*}
\end{proof}
\end{comment
\section*{Acknowledgements}
The first author would like to thank the University of Wollongong,
where part of this work was carried out.
The second author was partially supported by Alexander-von-Humboldt
fellowship 1137814 at the Otto-von-Guericke Universität Magdeburg and
by Australian Research Council Discovery Project grant DP120100097 at
the University of Wollongong. Further partial support toward both authors was provided by University of Wollongong Research Council 20124 Small Grant 22831024. They are grateful for their support.
\bibliographystyle{plain}
|
\section{Introduction}
Twitter a micro-blogging site, have become an integral part of the life of millions of users.
People use it to communicate with friends,family or acquaintances. With the increase in the number of users, demand for
analytics on this data is also growing. Twitter does not store gender, age, ethnicity or interests of a user.
These attributes of a user could be useful both for user experience as well as for consumption by brands in their social analytics.
\\
In this work we address the task of user gender classification, by leveraging observable
information such as the tweet behavior, linguistic content of the user's Twitter feed and the celebrities followed by the user.
\\
Main contributions of this work are following:
\begin{description}
\item[$\bullet$] It uses some new linguistic content based features, which were not used by state-of-the-art, particularly on Twitter data.
\item[$\bullet$] It describes completely new set of popular(celebrity) neighborhood based features such as \textit{age, gender and occupational area of the celebrity} which show promising results.
\item[$\bullet$] It reports that using combination of both features can perform better than various state-of-the-art techniques, and especially extraction of better popular network based features demands further investigation.
\end{description}
The paper is organized as follows. Section 2 introduces relevant and related work on user profiling for social
media, Twitter user attribute detection and topic models for
Twitter. Section 3 describes the datasets we have used or created. Section 4 we explains methodology we followed to clean and process the datasets, while Section 5
explains in detail use of various features and the intuition behind their usage. Section 6 describes results produced in different configurations. Finally, Section 7 draws final conclusions and outline future work.
\section{Background and Related work}
Inferring attributes of social media users is a growing area of interest, there have been many recent attempts to predict various hidden attributes of a user. Several of the recent works were focused on predicting ethnicity (Rao et al., 2011; Pennacchiotti and Popescu, 2011), age (Schler et al., 2006; Rosenthal and McKeown, 2011; Nguyen et al., 2011; Al Zamal et al., 2012), gender (Rao et al., 2010; Burger et al., 2011; Liu and Ruths, 2013; Al Zamal et al., 2012), Interests ( Lim and Datta, 2013), personality (Argamon et al., 2005; Schwartz et al., 2013) etc.\\
Most of the datasets prepared in these approaches were hand annotated, cherry picked or attributes are identified by the user itself. Also many of the datasets were limited to very broad attribute categories, such as age below 23 as attribute young and above 25 as old. Rao et al. (2011) used linguistic features to predict ethnicity and gender of facebook users. Using a very limited training data, they tried to evaluate fine grained ethnicity classes of Nigeria. Social or Network based features were first used by Pennacchiotti and Popescu (2011); they tried to predict if a Twitter user is African-American or not. \\
As preparing a good dataset is a big challenge, Schwartz et al. (2013) tried to explore this area and collected Facebook profiles labeled with personality type, gender, and age by administering a survey of users embedded in a personality test application.\\
Al Zamal et al. (2012) explored into social features by exploring homophily based features; such as using linguistic features of n-most-popular friends of a user on Twitter. Their final classifier obtains a \textbf{80.02\%} accuracy on binary gender classification (compared with the \textbf{85.67\%} accuracy this paper report below for same dataset).\\
Lim and Datta (2013) presented a novel approach for classifying user interests for e.g. Food, Politics, Charity etc. using celebrity(popular users) followed by a twitter users. \\
\section{Datasets}
In order to collect the datasets, it was necessary to decide on the two contrasting labels (for e.g ``male'' and ``female'') that would be used and identify users that could be reliably assigned one label or
the other. Even though we have used dataset which was already prepared by Zamal, still we had to manually inspect accounts collected
to confirm that the labeled assignments were correct.
\subsection{Gender Dataset}
This paper have used the same dataset prepared by Al Zamal et al. (2012) for gender prediction. Zamal's paper used a technique proposed in Mislove et al. (2011), which used users who had their full first and last names on Twitter; also their first name was one of the top 100 most common names on record with US social security department for baby boys/girls born in the year 2011. 131 male and 134 female labeled users were collected out of 192 each, because some users closed their Twitter account or locked their timeline for public access.
\\
\subsection{Popular neighborhood Dataset}
We have also created a dataset of the popular neighborhood or so called celebrity users followed by these 265 users. Here, we define celebrities(popular users) as users with more than 10,000 followers or who have been verified by Twitter itself, as was defined by Lim and Datta (2013). Thus, out of 108,619 distinct users followed by 265 users, we were left with 5,546 popular users. Once we had the list of these celebrity users we extracted various features which we would discuss in section 5.
\section{Methodology}
The initial results of applying user tweets based classification
did not seem very promising, a closer examination revealed
that the problem was in the quality of the Tweet text. Since the text
in this domain comes from Short messages, and the messages often contain various abbreviations, and are rife with
spelling errors. Thus, to deal with this, we implemented a process to auto-correct words that are greater than 3 letters in length as suggested by Prem Melville et al. (2013). This process leveraged a combination of Philips' metaphone algorithm and string-edit(TextBrew) distance for spell-correction; in addition to a general-purpose English dictionary, we used dictionary for Twitter short hands,
for e.g. \textit{``idk''} would be mapped to \textit{``I don't know''}. As seen on Twitter many people use a lot of camel-casing especially while using hashtags,
we tried to clean such camel-cases, for e.g. \textit{``\#BestDayEver''} would be cleaned as \textit{``Best Day Ever''}.
After cleaning tweets we saw 12\% gain in accuracy only based on user tweets based classification.
\begin{figure}[!htbp]
\centering
\includegraphics[width=80mm]{Methodology-AIR-Paper.pdf}
\caption{High level view of tweet based features extraction}
\label{overflow}
\end{figure}
\section{Features}
Previous works in this area have already explored some general lexicon features based on tweet text for e.g. K-top words used by each class.
In this paper we have not used these features, instead we used LIWC 2007 dictionary for extracting lexicon features. With LIWC based lexicon
features we got almost 3\% gain in accuracy compared to state-of-the-art lexicon features as shown in Section 6.
\\
We will now discuss all the features used in this paper for predicting gender of Twitter user.
\subsection{Tweet behavior based features}
Tweeting behavior can be a good parameter to distinguish between male and female, we have used almost all the features used by state-of-the-art.
\subsubsection{Tweet Frequency}
Let's say number of tweets extracted for a user be \textit{N} (in this paper we used N=1000) and
chronological difference in days between first and last tweet be \textit{C}, then ``tweet frequency'' is defined as
\begin{equation}
TF = \frac{N}{C}
\end{equation}
According to a study done by \textit{www.beevolve.com} on 36 million twitter users in 2012, it was found that on average female users tend to tweet more often than male counterparts.
\subsubsection{Hashtag Frequency}
Let ${T}_h$ be the total number of hashtags used by a user, then ``hashtag frequency'' is defined as
\begin{equation}
HF = \frac{{T}_h}{C}
\end{equation}
\subsubsection{Average Tweet Length}
``Average Tweet Length'' $TL$ represents the average length
of tweets. Here, $TL$ is defined as
\begin{equation}
TL= \frac{\sum \limits^{Recent N}_{i=0}Length(tweet_i)}{N}
\end{equation}
\subsubsection{Retweet Frequency}
Let ${T}_r$ be the total number of tweets retweeted by a user, then ``retweet frequency'' is defined as (Rao et al. 2010):
\begin{equation}
RF= \frac{{T}_r}{C}
\end{equation}
\subsubsection{Followers to following ratio}
The ratio between number of followers and number of friends has been used as a measure
of a user's tendency towards producing vs. consuming information on Twitter (Rao et al. 2010).
\subsubsection{Celebrity following tendency}
Let ${T}_p$ be the total number of celebrities or popular users followed by a user and ${T}_f$ be the total number , then `celebrity following tendency``
is defined as
\begin{equation}
CT= \frac{{T}_p}{C}
\end{equation}
\subsection{Linguistic content based features}
Linguistic content information encapsulates the main topics
of interest to the user as well as the user's lexical usage. In a study done by \textit{www.beevolve.com} (2012),
Female users talk more about family and fashion whereas the Twitter male users prefer technology, sports and entrepreneurship.
We explored a wide variety of linguistic content features using LIWC\footnote{Linguistic Inquiry and Word Count (LIWC) is a text analysis software program designed by James W. Pennebaker, Roger J. Booth, and Martha E. Francis. LIWC calculates the degree to which people use different categories of words across a wide array of texts, including emails, speeches, poems, or transcribed daily speech.} 2007 dictionary,
as detailed below.
\subsubsection{Count of \textit{I}, \textit{we}, \textit{you}, \textit{he-she} and \textit{they} words}
According to David Bamman et al. (2014), female Twitter user tend to use more personal diary writing style, where they might use more
\textit{I} and \textit{he-she} references.
\subsubsection{Parts of Speech}
\textbf{Pronouns} are generally associated with female authors (David Bamman et al., 2014), \textbf{Conjunction} such as \textit{and} were associated with female authors.
Other parts of speech such as articles, determiners or prepositions showed low gender association.
\subsubsection{Emotions and Emoticons}
Emotional terms such as sad, love, glad, etc. are more associated with female authors, while female author show more associativity with emotionally neutral sentences (David Bamman et al., 2014).
\\*
We have also used Emoticons such as :-), <3 etc. and their unicode counterparts. Emoticons tend to have more associativity with female authors.
\subsubsection{Categorical words}
Words related to Health, Money, Achievement, Society, Family etc. can be a good indicator for predicting gender of a twitter users.
As shown in many studies words related to ``Money and'' ``Finance'' are often used by male authors, while words related to ``Family'', ``Kinship'' and ``Society'' are more often
used by female authors on Twitter. Similarly ``Swear'' words are more often used by male authors rather than female authors (David Bamman et al., 2014; \textit{www.beevolve.com}, 2012).
In this papers we have used 50 such categories, each as one feature.
\begin{table}[!htbp]
\centering
\begin{tabular}{|c|c|}
\hline
\textit{\textbf{Categories}} & \textit{\textbf{Associativity}} \\ \hline
Pronouns & Female \\ \hline
Emotion terms & Female \\ \hline
Kinship terms & Female \\ \hline
CMC words (lol, omg) & Female \\ \hline
Numbers & Male \\ \hline
Technology words & Male \\ \hline
Swear words & Male \\ \hline
Assent & Female \\ \hline
Emoticons & Female \\ \hline
Hesitation & Female \\ \hline
\end{tabular}
\caption{Some of the categories and their known associativity with gender of user (David Bamman et al., 2014)}
\end{table}
\\
\\
In total we are using 64 different Linguistic content based features. These features proved to give better results than other simple Linguistic content based features
such as \textit{K-top} words.
\subsection{Popular neighborhood based features}
As we have already defined ``popular neighborhood'' or ``celebrity users'' as those users who are followed by any of the 265 users in our dataset having either (1) more than 10,000 followers or (2) are verified users\footnote{In
June 2008, Twitter launched a verification program, allowing celebrities, brands, businesses and public figures to get their accounts verified, there are currently approximately 92,000 Verified users on Twitter}.
Popular users (celebrities) use their actual name on Twitter and it is feasible to extract their features from websites like Wikipedia and Freebase. This gives us an opportunity to leverage
principle of homophilic association (in Twitter, a user
can exercise greater selectivity over who she follows than
who follows her) with greater accuracy in extracting neighborhood features.
\\
In the following subsections we would discuss our intuition behind each ``popular neighborhood based feature'' we have chosen and how we extracted those features.
\subsubsection{Age of celebrity}
We have used 5 categories(each as a feature) for a celebrity users age
\begin{description}
\item[$\bullet$] age < 23 years
\item[$\bullet$] 23<age<30
\item[$\bullet$] 30<age<40
\item[$\bullet$] 40<age<50
\item[$\bullet$] 50<age
\end{description}
for e.g. Justin Bieber would be of ``age<23'' category. The basic intuition behind taking celebrity age as feature for predicting gender is,
number of female users under age 30 years on Twitter outnumbered males under age 30, this gap even widens under age 20. Concept of homophilic association
suggests that users of same age category tend to follow each other more. Thus, probability of female following celebrities under age 23 years
would be more than male. We used Wikipedia to extract age of the celebrity.
Later in our experiment with this feature, we found that age of a celebrity has a weak associativity with the gender of its follower.
\subsubsection{Gender of celebrity}
It has been discussed by Bill Heil et al. (2009), that an average man is almost twice more likely to follow another man than a woman.
We have used ``male'' and ``female'' category of celebrity as feature(for e.g. Sachin Tendulkar would be categorized as male), and ignored gender neutral Twitter users such as Brand and Business accounts.
\begin{table}[h]
\centering
\begin{tabular}{|c|c|c|}
\hline
& \textit{a female} & \textit{a male} \\ \hline
Female follows & 44\% & 56\% \\ \hline
Male follows & 35\% & 65\% \\ \hline
\end{tabular}
\caption{Tendency of male and female following other male and female (Bill Heil et al., 2009)}
\end{table}
\\
Since it is almost impossible to find gender of the complete neighborhood of a user, we have used just popular neighborhood.
We extracted gender of a celebrity using Wikipedia by counting \textit{he, she, her, his, him etc.} references on that celebrities page.
In our experiment we found promising evidence suggesting gender of neighborhood could be helpful in predicting gender of user itself.
\subsubsection{Celebrities ``famous for''}
Work of Lim and Datta (2012) suggests that celebrities represent an interest category, and they leveraged these interests to discover communities who follow specific interests on Twitter.
In this paper we have also used similar categories to find interests of a user using the kind of celebrity they follow. We have used followed following categories each as one feature.
\\
\begin{table}[!htbp]
\centering
\begin{tabular}{|c|}
\hline
\textbf{``famous for'' Categories of celebrities}
\\ \hline
\begin{tabular}{l l}
$\bullet$ Acting & $\bullet$ Art \\
$\bullet$ Entertainment & $\bullet$ Entrepreneur \\
$\bullet$ Writing & $\bullet$ Music \\
$\bullet$ Politics & $\bullet$ Religious \\
$\bullet$ Science \& Technology & $\bullet$ Security \\
$\bullet$ Social & $\bullet$ Sports \\
$\bullet$ Miscellaneous & \\
\end{tabular}
\\ \hline
\end{tabular}
\end{table}
\\
According to a study conducted by \textit{www.beevolve.com} on 36 million twitter users, it was found that some of the broad categories were more
distinctive then other. And these are of interests could be used predict gender and other attributes of a user.
\\
\begin{figure}[!htbp]
\centering
\includegraphics[width=85mm]{interest.png}
\caption{Gender distribution by top 10 categories of interest on Twitter (\textit{www.beevolve.com}, 2012)}
\label{overflow}
\end{figure}
\\
\\
\\
\\
\\
\\
\\
\section{Results}
We would evaluate our results in terms of overall accuracy, A 10-fold cross-validation
was done to assess the performance of the SVM-based classifiers at inferring the gender of a Twitter user.
\\*
We will discuss the results in following three stages:
\subsection{Using tweet behavior and linguistic features}
\begin{table}[!htbp]
\centering
\begin{tabular}{|c|c|}
\hline
\textit{\textbf{Configuration}} & \textit{\textbf{Accuracy}} \\ \hline
LIWC only(with last 200 tweets) & 77.95\% \\ \hline
LIWC only(with last 1000 tweets) & 80.43\% \\ \hline
LIWC + Tweet Behavior features & \textbf{82.26\%} \\ \hline
\end{tabular}
\caption{The overall accuracy while using all the tweet based features which excludes homophily features}
\end{table}
\subsection{Using popular neighborhood features}
\begin{table}[!htbp]
\centering
\begin{tabular}{|c|c|}
\hline
\textit{\textbf{Configuration}} & \textit{\textbf{Accuracy}} \\ \hline
Age of celebrity & 56.05\% \\ \hline
Celebrity ``famous for'' & 67.23\% \\ \hline
Gender of celebrity & \textbf{71.57\%} \\ \hline
Combined & 66.84\% \\ \hline
\end{tabular}
\caption{The overall accuracy using popular neighborhood features}
\end{table}
\subsection{Using all the features combined}
\begin{table}[!htbp]
\centering
\begin{tabular}{|c|c|c|l|}
\hline
& \textit{True Male} & \textit{True Female} & \textbf{Precision} \\ \hline
Pred. Male & 115 & 17 & 87.12\%\ \\ \hline
Pred. Female & 21 & 112 & 84.21\% \\ \hline
\textbf{Recall} & 84.56\% & 86.82\% & \\ \hline
\end{tabular}
Overall accuracy = \textbf{85.67\%}
\caption{The final confusion matrix and overall accuracy using all the features}
\end{table}
\section{Conclusion and future}
In this paper, we evaluated the extent to which features
present in a Twitter user’s popular neighbors can improve
the inference of attributes possessed by the user herself.
Our results support several noteworthy conclusions, which we discuss here.
\subsection{Usefulness of linguistic features}
Liguistic features such as associativity of a particular category of words towards the gender of a user as suggested by David Bamman et al. (2014), produced satisfactory results.
\\*
\subsection{Popular neighborhood is useful and feature rich}
Pennacchiotti et. al (2011), discussed in their paper that users following more celebrities tend to give better results in their homophily(network) based approach. We verified in this paper that
popular neighborhood can be leveraged to extract various rich and accurate features using already available knowledge bases.\\
We have observed that gender of a celebrity can be a good feature, and proves the study conducted by Bill Heil et al. (2009), which suggests tendency of males to follow male is twice than following female.\\
Information regarding popular user's domain of work can be useful for determining users interest which in turn can be useful in gender classification.
\\ \\
In this paper we have not used other homophily features such as linguistic features of popular neighborhood,
it would interesting to see combined results of homophily and celebrity based features used in this paper. We can also explore the possibility of using these features to infer other user attributes such as ``age'' and ``political orientation''
\section{Acknowledgments}
I would like to thank Rahul Tejwani for his help and active participation in various discussions for this work.
I would also like to thank Professor Rohini Srihari and all those who helped me throughout this project.
\bibliographystyle{abbrv}
|
\section{Introduction}
The inverse problem in electrical impedance tomography (EIT) consists
of reconstructing an electrical conductivity distribution in the
interior of an object from electro-static boundary measurements on the
surface of the object. EIT is an emerging technology with applications
in medical imaging \cite{Holder2005}, geophysics \cite{Abubakar2009} and
industrial tomography \cite{York2001}. The underlying mathematical problem is
known as the Calder\'on problem in recognition of Calder\'on's seminal
paper \cite{Calderon1980}.
Consider a bounded domain $\Omega\subset \mathbb{R}^n,\; n\geq 2, $ with smooth boundary $\po.$ In
order to consider partial boundary measurements we introduce the
subsets $\gamN,\gamD \subseteq \po$ for the Neumann and Dirichlet data
respectively. Let $\sigma \in L^\infty(\Omega)$ with $0< c\leq \sigma$ a.e.
denote the conductivity distribution in $\Omega$. Applying a boundary
current flux $g$ (Neumann condition) through $\gamN\subseteq \po$
gives rise to the interior electric potential $u$ characterized as the
solution to
\begin{equation}
\nabla \cdot(\sigma\nabla u) = 0 \text{ in } \Omega,\quad \sigma \frac{\partial u}{\partial \nu} = g \text{ on } \partial\Omega, \quad \int_{\gamD} u|_{\po}\,ds = 0, \label{pde}
\end{equation}
where $\nu$ is an outward unit normal to $\po.$ The latter condition
in \eref{pde} is a grounding of the total electric potential along the
subset $\gamD\subseteq \po.$ To be precise we define the spaces
\begin{align*}
L_\diamond^2(\po) &\equiv \{g \in L^2(\partial\Omega) \mid \int_{\po}
g \,ds = 0\}, \\
\hmhalf &\equiv \{g \in H^{-1/2}(\partial\Omega)
\mid \inner{g,1} = 0\},
\end{align*}
consisting of boundary functions with mean zero, and the spaces
\begin{align*}
H_{\gamD}^1(\Omega) &\equiv \{u\in H^1(\Omega)\mid u|_{\po} \in
H_{\gamD}^{1/2}(\po)\,\}, \\
H_{\gamD}^{1/2}(\po) &\equiv \{f\in H^{1/2}(\po)\mid \int_{\gamD} f\,ds=0\},
\end{align*}
consisting of functions with mean zero on $\gamD$ designed to
encompass the partial boundary data. Using standard elliptic theory
it follows that \eref{pde} has a unique solution $u \in
H_{\gamD}^1(\Omega)$ for any $g \in \hmhalf$. This defines the
Neumann-to-Dirichlet map (ND-map) $\Lambda_\sigma$ as an operator from
$H_\diamond^{-1/2}(\po)$ into $H_{\gamD}^{1/2}(\po)$ by $g \mapsto
u|_{\po}$, and the partial ND-map as $g \mapsto (\Lambda_\sigma
g)|_{\gamD}$.
The data for the classical Calder\'on problem is the full operator
$\Lambda_{\sigma}$ with $\gamD = \gamN = \partial\Omega.$
The problem is well-studied and there are numerous publications addressing different
aspects of its solution; we mention only a few: the uniqueness and
reconstruction problem was solved in \cite{SylvesterUhlmann1987,
Nachman1988, Novikov1988, Nachman1996, AstalaPaivarinta2006,
HabermanTataru2013} using the so called complex geometrical optics (CGO)
solutions; for a recent survey see \cite{Uhlmann2009}. Stability
estimates of log type were obtained in \cite{Alessandrini1988,
Alessandrini1990} and shown to be optimal in
\cite{Mandache2001}. Thus any computational algorithm must rely on
regularization. Such computational regularization algorithms
following the CGO approach were designed, implemented and analysed in
\cite{SiltanenMuellerIsaacson2001,KnudsenLassasMuellerSiltanen2009,
BikowskiKnudsenMueller2011,DelbaryHansenKnudsen2012,DelbaryKnudsen2014}.
Recently the partial data Calder\'on problem has been studied
intensively. In 3D uniqueness has been proved under certain conditions
on $\gamD$ and $\gamN$
\cite{BukhgeimUhlmann2002,KenigSjostrandUhlmann2007,Knudsen2006,Zhang2012,Isakov2007}, and
in 2D the general problem with localized data i.e. $\gamD = \gamN =
\Gamma$ for some, possibly small, subset $\Gamma\subseteq\po$ has been
shown to posses uniqueness \cite{ImanuvilovUhlmannYamamoto2010}. Also
stability estimates of log-log type have been obtained for the partial
problem \cite{HeckWang2006}; this suggests that the partial data
problem is even more ill-posed and hence requires more regularization
than the full data problem. Recently a computational algorithm for the
partial data problem in 2D was suggested and investigated in
\cite{HamiltonSiltanen}.
A general approach to linear inverse problems with sparsity
regularization was given in \cite{daubechies2004}, and in
\cite{bredies2009,bonesky2007} the method
was adapted to non-linear problems using a so-called generalized conditional
gradient method. In \cite{gehre2012,jin2011,jin2012} the method was applied to the reconstruction problem in EIT with full boundary
data. For other approaches to EIT using optimization methods we refer
to \cite{Borcea2002}.
In this paper we will focus on the partial data problem for which we
develop a reconstruction algorithm based on a least squares
formulation with sparsity regularization. The results are twofold:
first we extend the full data algorithm of \cite{jin2012} to the case
of partial data, second we show how prior information about the spatial
location of the perturbation in the conductivity can be used in the
design of a spatially varying regularization parameter. We will restrict the treatment to 2D, however everything extends to 3D with some minor assumptions on the regularity of the Neumann data \cite{GardeKnudsen2014}.
The data considered here consist of a finite
number of Cauchy data taken on the subsets $\gamD$ and $\gamN,$ i.e.
\begin{equation}
\{(f_k,g_k) \mid g_k \in \hmhalf, \; \supp(g_k)\subseteq \gamN, f_k = \Lambda_{\sigma} g_k|_{\gamD} \}_{k=1}^K,\; K \in \mathbb{N}. \label{data}
\end{equation}
We assume that the true conductivity is given as $\sigma = \sigma_0 + \delta\sigma$, where $\sigma_0$ is a known background conductivity. Define the closed and convex subset
\begin{equation}
\arum_0 \equiv \{\delta\gamma \in H_0^1(\Omega)\mid c\leq \sigma_0+\delta\gamma\leq c^{-1} \text{ a.e. in } \Omega\} \label{a0ref}
\end{equation}
for some $c \in (0,1)$, and $\sigma_0\in H^1(\Omega)$ where $c\leq \sigma_0\leq c^{-1}$. Similarly define
\[
\arum\equiv \arum_0 + \sigma_0 = \{\gamma\in H^1(\Omega)\mid c\leq \gamma\leq c^{-1} \text{ a.e. in } \Omega, \gamma|_{\po} = \sigma_0|_{\po}\}.
\]
The inverse problem is then to approximate $\delta\sigma\in\arum_0$ given the data \eref{data}.
Let $\{\psi_j\}$ denote a chosen orthonormal basis for $H_0^1(\Omega).$ For sparsity regularization we approximate $\delta\sigma$ by $\argmin_{\delta\gamma\in\arum_0}\Psi(\delta\gamma)$ using the following Tikhonov functional \cite{jin2012}
\begin{equation}
\Psi(\delta\gamma) \equiv \sum_{k=1}^K R_k(\delta\gamma) + P(\delta\gamma), \enskip\delta\gamma\in\arum_0, \label{psieq}
\end{equation}
with
\[
R_k(\delta\gamma) \equiv \frac{1}{2}\normm{\Lambda_{\sigma_0+\delta\gamma}g_k - f_k}_{L^2(\gamD)}^2, \quad P(\delta\gamma) \equiv \sum_{j=1}^\infty \alpha_j\absm{c_j},
\]
for $c_j \equiv \inner{\delta\gamma,\psi_j}$. The regularization parameter $\alpha_j>0$ for the
sparsity-promoting $\ell_1$ penalty term $P$ is distributed such that
each basis coefficient can be regularized differently; we will return to
this in \sref{sec:prior}. It should be noted how easy and natural the
use of partial data is introduced in this way, simply by only
minimizing the discrepancy on $\gamD$ on which the Dirichlet data is known and ignoring the rest of the boundary.
This paper is organised as follows: in \sref{sec:SparseReconstruction}
we derive the Fr\'echet derivative of $R_k$ and reformulate the
optimization problem using the generalized conditional gradient method as a
sequence of linearized optimization problems. In \sref{sec:prior} we
explain the idea of the spatially dependent regularization parameter
designed for the use of prior information. Then in
\sref{sec:NumericalResults} we show the feasibility of the algorithm
by several numerical examples, and finally we conclude in
\sref{sec:conclusions}.
\section{Sparse Reconstruction}\label{sec:SparseReconstruction}
In this section the sparse reconstruction of $\dsr$ based on the optimization problem \eqref{psieq}, is investigated for a bounded domain $\Omega\subset\rum{R}^2$ with smooth boundary $\partial\Omega$. The penalty term emphasizes that $\dsr$ should only be expanded by few basis functions in a given orthonormal basis. Using a distributed regularization parameter, it is
possible to further apply prior information about which basis functions that should be included in the expansion of $\dsr$. The partial data problem comes into play in the discrepancy term, in which we only fit the data on part of the boundary. Ultimately, this leads to the algorithm given in Algorithm \ref{alg1} at the end of this section.
Denote by $F_g(\sigma)$ the unique solution to \eref{pde} and let $\fob_g(\sigma)$ be its trace (note that $\Lambda_\sigma g = \fob_g(\sigma)$). Let $\gamma\in\arum$, $g\in L^p(\po)\cap\hmhalf$ for $p>1$, then following the proofs of Theorem 2.2 and Corollary 2.1 in \cite{jin2011} whilst applying the partial boundary $\gamD$ we have
\begin{equation}
\lim_{\substack{\normm{\ds}_{H^1(\Omega)}\to 0 \\ \gamma+\ds\in\arum}}\frac{\normm{\fob_g(\gamma+\ds)-\fob_g(\gamma) - (\fob_g)'_\gamma\ds}_{H^{1/2}_{\gamD}( \partial\Omega)}}{\normm{\ds}_{H^1(\Omega)}} = 0. \label{fpplimit}
\end{equation}
Here $(\fob_g)'_\gamma$ is the linear map, that maps $\eta$ to $w|_{\po}$, where $w$ is the unique solution to
\begin{equation}
-\nabla\cdot(\gamma\nabla w) = \nabla\cdot(\ds\nabla F_g(\gamma)) \text{ in } \Omega, \quad \sigma\frac{\partial w}{\partial\nu} = 0 \text{ on } \partial\Omega, \quad \int_{\gamD}w|_{\partial\Omega}\,ds = 0. \label{fprimepde}
\end{equation}
It is noted that $(\fob_g)'_\gamma$ resembles a Fr\'echet derivative of $\fob_g$ evaluated at $\gamma$ due to \eref{fpplimit}, however $\arum$ is not a linear vector space, thus the requirement $\gamma,\gamma+\ds\in\arum$.
The first step in minimizing $\Psi$ using a gradient descent type iterative algorithm is to determine a derivative to the discrepancy terms $R_k$.
\begin{lemma}
Let $\gamma=\sigma_0+\delta\gamma$ for $\delta\gamma\in\arum_0$, and $\chi_{\gamD}$ be a characteristic function on $\gamD$. Then
\begin{equation}
G_k \equiv -\nabla F_{g_k}(\gamma)\cdot\nabla F_{\chi_{\gamD}(\Lambda_\gamma g_k-f_k)}(\gamma)\in L^r(\Omega)\subset H^{-1}(\Omega) \label{nablaJ}
\end{equation}
for some $r>1$, and the Fr\'echet derivative $(R_k)'_{\delta\gamma}$ of $R_k$ on $H_0^1(\Omega)$ evaluated at $\delta\gamma$ is given by
\begin{equation}
(R_k)'_{\delta\gamma}\eta = \int_{\Omega} G_k\ds \,dx, \enskip \delta\gamma+\ds\in \arum_0. \label{Jprimeeq}
\end{equation}
\end{lemma}
\begin{proof}
For the proof the index $k$ is suppressed. First it is proved that $G\in L^r(\Omega)$ for some $r>1$, which is shown by estimates on $F_g(\gamma)$ and $F_h(\gamma)$ where $h \equiv \chi_{\gamD}(\Lambda_\gamma g-f)$. Note that $\Lambda_\gamma g \in H_{\gamD}^{1/2}(\po)$ and $f\in L^2_\diamond(\gamD)$, i.e. $h \in L^2_\diamond(\po) \subset L^2(\po)\cap H_\diamond^{-1/2}(\po)$. Now using \cite[Theorem 3.1]{jin2011}, there exists $Q>2$ such that
\begin{equation}
\normm{F_h(\gamma)}_{W^{1,q}(\Omega)} \leq C\normm{h}_{L^2(\po)}, \label{Fweq}
\end{equation}
where $q\in(2,Q)\cap [2,4]$. Since $F_g(\gamma)\in H_{\gamD}^1(\Omega)$ then $\absm{\nabla F_g(\gamma)} \in L^2(\Omega)$. It has already been established in \eref{Fweq} that $F_h(\gamma) \in W^{1,q}(\Omega)$ for $q\in(2,\min\{Q,4\})$, so $\absm{\nabla F_h(\gamma)}\in L^q(\Omega)$. By H{\"o}lder's generalized inequality
\[
G = -\nabla F_g(\gamma)\cdot \nabla F_h(\gamma) \in L^r(\Omega), \enskip \tfrac{1}{r} = \tfrac{1}{2} + \tfrac{1}{q},
\]
and as $q > 2$ then $r > 1$. Let $r'$ be the conjugate exponent to $r$, then $r'\in[1,\infty)$, i.e. the Sobolev imbedding theorem \cite{sobolev} implies that $H^1(\Omega)\imbed L^{r'}(\Omega)$ as $\Omega\subset\rum{R}^2$. Thus $G \in (L^{r'}(\Omega))'\subset (H^1(\Omega))' \subset (H_0^1(\Omega))' = H^{-1}(\Omega)$.
Now it will be shown that $R'_{\delta\gamma}$ can be identified with $G$. $R'_{\delta\gamma}\eta$ is by the chain rule (utilizing that $\Lambda_\gamma g = \fob_g(\gamma)$) given as
\begin{equation}
R'_{\delta\gamma}\eta = \int_\po \chi_{\gamD}(\Lambda_\gamma g-f)(\fob_{g})_\gamma'\ds \,ds, \label{expectedprime}
\end{equation}
where $\chi_{\gamD}$ is enforcing that the integral is over $\gamD$. The weak formulations of \eref{pde}, with Neumann data $\chi_{\gamD}(\Lambda_\gamma g-f)$, and \eref{fprimepde} are
\begin{align}
\int_{\Omega} \gamma \nabla F_{\chi_{\gamD}(\Lambda_\gamma g-f)}(\gamma)\cdot \nabla v\,dx &= \int_{\po} \chi_{\gamD}(\Lambda_\gamma g-f)v|_{\po}\,ds, \enskip \forall v\in H^1(\Omega), \label{weak1} \\
\int_{\Omega} \gamma \nabla w\cdot \nabla v\,dx &= -\int_{\Omega} \ds\nabla F_{g}(\gamma)\cdot \nabla v\,dx, \enskip \forall v\in H^1(\Omega). \label{weak2}
\end{align}
Now by letting $v \equiv w$ in \eref{weak1} and $v \equiv F_{\chi_{\gamD}(\Lambda_\gamma g-f)}(\gamma)$ in \eref{weak2}, we obtain using the definition $w|_{\po} = (\fob_g)'_\gamma\eta$ that
\begin{align*}
R'_{\delta\gamma}\ds &= \int_\po \chi_{\gamD}(\Lambda_\gamma g-f)(\fob_{g})_\gamma'\ds \,ds = \int_{\Omega} \gamma \nabla F_{\chi_{\gamD}(\Lambda_\gamma g-f)}(\gamma)\cdot \nabla w\,dx \\
&= -\int_{\Omega} \ds\nabla F_{g}(\gamma)\cdot \nabla F_{\chi_{\gamD}(\Lambda_\gamma g-f)}(\gamma)\,dx = \int_{\Omega}G\ds \,dx.
\end{align*}
\end{proof}
\begin{remark}
It should be noted that $(R_k)'_{\delta\gamma}$ is related to the Fr\'echet derivative $\Lambda_\gamma'$ of $\gamma\mapsto \Lambda_\gamma$ evaluated at $\gamma$, by $(R_k)'_{\delta\gamma}\eta = \int_{\gamD}(\Lambda_\gamma g_k-f_k)\Lambda_\gamma'[\eta]g_k ds$.
\end{remark}
Define
\[
R'_{\delta\gamma} \equiv \sum_{k=1}^K (R_k)'_{\delta\gamma} = -\sum_{k=1}^K\nabla F_{g_k}(\gamma)\cdot\nabla F_{\chi_{\gamD}(\Lambda_{\gamma} g_k-f_k)}(\gamma).
\]
For a gradient type descent method, we seek to find a direction $\eta$ for which the discrepancy decreases. As $R_{\delta\gamma}'\in H^{-1}(\Omega)$ it is known from Riesz' representation theorem that there exists a unique function in $H_0^1(\Omega)$, denoted by $\nabla_s R(\delta\gamma)$, such that
\begin{equation}
R_{\delta\gamma}'\eta = \inner{\nabla_s R(\delta\gamma),\eta}_{H^1(\Omega)},\enskip \eta\in H_0^1(\Omega). \label{sobograd}
\end{equation}
Now $\eta \equiv -\nabla_s R(\delta\gamma)$ points in the steepest descend direction among the viable directions. Furthermore, since $\nabla_s R(\delta\gamma)|_{\po} = 0$ the boundary condition $\delta\sigma|_{\po} = 0$ will automatically be fulfilled for the approximation. In \cite{sobolevgradient} $\nabla_s R(\delta\gamma)$ is called a Sobolev-gradient, and it is the unique solution to
\[
(-\Delta+1)v=R_{\delta\gamma}' \text{ in } \Omega, \quad v = 0\text{ on } \po,
\]
for which \eref{sobograd} is the weak formulation. In each iteration step we need to determine a step size $s_i$ for an algorithm resembling a steepest descent $\delta\gamma_{i+1} = \delta\gamma_i - s_i\nabla_sR(\delta\gamma_i)$. Here a Barzilai-Borwein step size rule \cite{BBrules,sparsa,jin2012} will be applied, for which we determine $s_i$ such that $\frac{1}{s_i}(\delta\gamma_i-\delta\gamma_{i-1}) = \frac{1}{s_i}(\gamma_i-\gamma_{i-1})\simeq \nabla_s R(\delta\gamma_i)-\nabla_s R(\delta\gamma_{i-1})$ in the least-squares sense
\begin{equation}
s_i \equiv \argmin_{s} \normm{s^{-1}(\delta\gamma_i-\delta\gamma_{i-1})-(\nabla_s R(\delta\gamma_i)-\nabla_s R(\delta\gamma_{i-1}))}_{H^1(\Omega)}^2. \label{bbrule}
\end{equation}
Assuming that $\innerm{\delta\gamma_i-\delta\gamma_{i-1},\nabla_s R(\delta\gamma_i)-\nabla R(\delta\gamma_{i-1})}_{H^1(\Omega)}\neq 0$ yields
\begin{equation}
s_i = \frac{\normm{\delta\gamma_i-\delta\gamma_{i-1}}_{H^1(\Omega)}^2}{\innerm{\delta\gamma_i-\delta\gamma_{i-1},\nabla_s R(\delta\gamma_i)-\nabla_sR(\delta\gamma_{i-1})}_{H^1(\Omega)}}. \label{bbstepsize}
\end{equation}
A maximum step size $s_{\max}$ is enforced to avoid the situations where $\innerm{\delta\gamma_i-\delta\gamma_{i-1},\nabla_s R(\delta\gamma_i)-\nabla R(\delta\gamma_{i-1})}_{H^1(\Omega)}\simeq 0$.
With inspiration from \cite{sparsa}, $s_i$ will be initialized by \eref{bbstepsize}, after which it is thresholded to lie in $[s_{\min},s_{\max}]$, for positive constants $s_{\min}$ and $s_{\max}$. It is noted in \cite{sparsa} that Barzilai-Borwein type step rules lead to faster convergence if we do not restrict $\Psi$ to decrease in every iteration. Allowing an occasional increase in $\Psi$ can be used to avoid places where the method has to take many small steps to ensure the decrease of $\Psi$. Therefore, one makes sure that the following so called weak monotonicity is satisfied, which compares $\Psi(\delta\gamma_{i+1})$ with the most recent $M$ steps. Let $\tau\in(0,1)$ and $M\in\rum{N}$, then $s_i$ is said to satisfy the weak monotonicity with respect to $M$ and $\tau$ if the following is satisfied \cite{sparsa}
\begin{equation}
\Psi(\delta\gamma_{i+1}) \leq \max_{i-M+1\leq j \leq i}\Psi(\delta\gamma_j) - \frac{\tau}{2s_i}\normm{\delta\gamma_{i+1}-\delta\gamma_i}_{H^1(\Omega)}^2.\label{weakmono}
\end{equation}
If \eref{weakmono} is not satisfied, the step size $s_i$ is reduced until this is the case. To solve the non-linear minimization problem for \eqref{psieq} we iteratively solve the following linearized problem
\begin{align}
\zeta_{i+1} &\equiv \argmin_{\delta\gamma\in H_0^1(\Omega)}\left[\frac{1}{2}\normm{\delta\gamma - (\delta\gamma_i-s_i\nabla_sR(\delta\gamma_i))}_{H^1(\Omega)}^2 + s_i\sum_{j=1}^\infty \alpha_j\absm{c_j}\right], \label{upsiloneq} \\
\delta\gamma_{i+1} &\equiv \proja(\zeta_{i+1}). \notag
\end{align}
Here $\{\psi_j\}$ is an orthonormal basis for $H_0^1(\Omega)$ in the $H^1$-metric, and $\proja$ is a projection of $H_0^1(\Omega)$ onto $\arum_0$ to ensure that \eqref{pde} is solvable (note that $H_0^1(\Omega)$ does not imbed into $L^\infty(\Omega)$, i.e. $\zeta_{i+1}$ may be unbounded). By use of the map $\srum_\beta:\rum{R}\to\rum{R}$ defined below, known as the soft shrinkage/thresholding map with threshold $\beta > 0$,
\begin{equation}
\ssc{\beta}(x) \equiv \sign(x)\max\{\absm{x}-\beta,0\},\enskip x\in\rum{R}, \label{softoperator}
\end{equation}
the solution to \eref{upsiloneq} is easy to find directly (see also \cite[Section 1.5]{daubechies2004}):
\begin{equation}
\zeta_{i+1} = \sum_{j=1}^\infty \ssc{s_i\alpha_j}(d_j)\psi_j, \label{softstep}
\end{equation}
where $d_j\equiv\innerm{\delta\gamma_i-s_i\nabla_sR(\delta\gamma_i),\psi_j}_{H^1(\Omega)}$ are the basis coefficients for $\delta\gamma_i-s_i\nabla_sR(\delta\gamma_i)$.
The projection $\proja : H_0^1(\Omega)\to \arum_0$ is defined as
\begin{equation*}
\proja(v) \equiv T_c(\sigma_0 + v) - \sigma_0, \enskip v\in H_0^1(\Omega),
\end{equation*}
where $T_c$ is the following truncation that depends on the constant $c\in(0,1)$ in \eref{a0ref}
\begin{equation*}
T_c(v) \equiv \begin{cases}
c & \text{where } v < c \text{ a.e.}, \\
c^{-1} & \text{where } v > c^{-1} \text{ a.e.}, \\
v & \text{else.}
\end{cases}
\end{equation*}
Since $\sigma_0\in H^1(\Omega)$ and $c\leq \sigma_0\leq c^{-1}$, it follows directly from \cite[Lemma 1.2]{Stampacchia_1965} that $T_c$ and $\proja$ are well-defined, and it is easy to see that $\proja$ is a projection. It should also be noted that $0\in \arum_0$ since $c\leq \sigma_0\leq c^{-1}$, thus we may choose $\delta\gamma_0 \equiv 0$ as the initial guess in the algorithm.
The algorithm is summarized in Algorithm \ref{alg1}. In this paper the stopping criteria is when the step size $s_i$ gets below a threshold $s_\text{stop}$.
\begin{remark}
Note that $\sum_j \innerm{\delta\gamma_i-s_i\nabla_sR(\delta\gamma_i) , \psi_j}_{H^1(\Omega)} \psi_j$ corresponds to only having the discrepancy term in \eref{upsiloneq}, while the penalty term corresponds to changing these coefficients with the soft thresholding.
\end{remark}
\begin{remark}
The non-linearity of $\gamma\mapsto\Lambda_\gamma$ leads to a non-convex discrepancy term, i.e. $\Psi$ is non-convex. So the best we can hope is to find a local minimum.
\end{remark}
\begin{algorithm} \caption{Sparse Reconstruction for Partial Data EIT} \label{alg1}
\begin{algorithmic}
\State Set $\delta\gamma_0 := 0$.
\While{stopping criteria not reached}
\State Set $\gamma_i := \sigma_0 + \delta\gamma_i$.
\State Compute $\Psi(\delta\gamma_i)$.
\State Compute $R_{\delta\gamma_i}' := -\sum_{k=1}^K\nabla F_{g_k}(\gamma_i)\cdot\nabla F_{\chi_{\gamD}(\Lambda_{\gamma_i} g_k-f_k)}(\gamma_i)$.
\State Compute $\nabla_s R(\delta\gamma_i)\in H_0^1(\Omega)$ such that $R_{\delta\gamma_i}'\ds = \innerm{\nabla_s R(\delta\gamma_i),\ds}_{H^1(\Omega)}$.
\State Compute step length $s_i$ by \eref{bbstepsize}, and decrease it till \eref{weakmono} is satisfied.
\State Compute the basis coefficients $\{d_j\}_{j=1}^\infty$ for $\delta\gamma_i-s_i\nabla_sR(\delta\gamma_i)$.
\State Update $\delta\gamma_{i+1} := \proja\para{\sum_{j=1}^\infty \ssc{s_i\alpha_j}(d_j)\psi_j}$.
\EndWhile
\State Return final iterate of $\delta\gamma$.
\end{algorithmic}
\end{algorithm}
\begin{remark}
The main computational cost lies in computing $R_{\delta\gamma_i}'$, which involves solving $2K$ well-posed PDE's (note that $F_{g_k}(\gamma_i)$ can be reused from the evaluation of $\Psi$). It should be noted that each of the $2K$ problems consists of solving the same problem, but with different boundary conditions, which leads to only having to assemble and factorize the FEM matrix once per iteration.
\end{remark}
\section{Prior Information} \label{sec:prior}
Prior information is typically introduced in the penalty term $P$ for Tikhonov-like functionals, and here the regularization parameter determines how much this prior information is enforced. In the case of sparsity regularization this implies knowledge of how sparse we expect the solution is in general. Instead of applying the same prior information for each basis function, a distributed parameter is applied. Let
\[
\alpha_j \equiv \alpha\mu_j,
\]
where $\alpha$ is a usual regularization parameter, corresponding to the case where no prior information is considered about specific basis functions. The $\mu_j\in(0,1]$ will be used to weigh the penalty depending on whether a specific basis function should be included in the expansion of $\dsr$. The $\mu_j$ are chosen as
\[
\mu_j = \begin{cases}
1, \quad & \text{no prior on $c_j$}, \\ \sim 0, & \text{prior that $c_j\neq 0$},
\end{cases}
\]
i.e. if we know that a coefficient in the expansion of $\dsr$ should be non-zero, we can choose to penalize that coefficient less.
\subsection{Applying the FEM basis}
In order to improve the sparsity solution for finding small
inclusions, it seems appropriate to include prior information about
the support of the inclusions. There are different methods available
for obtaining such information assuming piecewise constant
conductivity \cite{kirsch2007,HarrachUllrich2013} or real analytic
conductivity \cite{HarrachSeo2010}. An example of the reconstruction of $\supp\delta\sigma$ is shown in \fref{fig:monomethod}, where it is observed that numerically it is possible to reconstruct a reasonable convex approximation to the support. Thus, it is possible to acquire estimates of $\supp \delta\sigma$ \emph{for free}, in the sense that it is gained directly from the data without further assumptions.
\begin{figure}[h
\centering
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig1a}
\caption{{}} \label{fig:kitephantom_a}
\end{subfigure
\hspace{2cm}
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig1b}
\caption{{}} \label{fig:mono}
\end{subfigure
\caption{\textbf{(a):} Phantom with kite-shaped piecewise constant inclusion $\delta\sigma$. \textbf{(b):} Reconstruction of $\supp \delta\sigma$ using monotonicity relations from the approach in \cite{HarrachUllrich2013} by use of simulated noiseless data.}
\label{fig:monomethod}
\end{figure}
Another approach is to consider
other reconstruction methods such as total variation regularization
that tends to give good approximations to the support, but has issues
with reconstructing the contrast if the amplitude of $\dsr$ is
large as seen in \sref{sec:comparison}. The idea is to be able to apply such information in the
sparsity algorithm in order to get good contrast reconstruction while
maintaining the correct support, even for the partial data problem.
Suppose that as a basis we consider a finite element method (FEM)
basis $\{\psi_j\}_{j=1}^N$ for the subspace $V_h\subseteq H_0^1(\Omega)$ of piecewise affine
elements. This basis comprises basis functions that are piecewise
affine with degrees of freedom at the mesh nodes,
i.e.\ $\psi_j(x_k) = \delta_{j,k}$ at mesh node $x_k$ in the applied
mesh. Let $\dsr\in V_h$, then $\dsr(x) = \sum_j \dsr(x_j)\psi_j(x)$,
i.e.\ for each node there is a basis function for which the coefficient
contains local information about the expanded function; this is
convenient when applying prior information about the support of an
inclusion. Note that the FEM basis functions are not mutually orthogonal, since basis functions corresponding to neighbouring nodes are non-negative and have
overlapping support. However, for any non-neighbouring pair of nodes
the corresponding basis functions are orthogonal.
When applying the FEM basis for mesh nodes $\{x_j\}_{j=1}^N$, the corresponding functional is
\[
\Psi(\delta\gamma) = \frac{1}{2}\sum_{k=1}^K \normm{\Lambda_{\sigma_0+\delta\gamma}g_k - f_k}_{L^2(\gamD)}^2 + \sum_{j=1}^N \alpha_j\absm{\delta\gamma(x_j)}.
\]
It is evident that the penalty corresponds to determining inclusions with small support, and prior information on the sparsity corresponds to prior information on the support of $\dsr$. We cannot directly utilize \eref{softstep} due to the FEM basis not being an orthonormal basis for $H_0^1(\Omega)$, and instead we suggest the following iteration step:
\begin{align}
\zeta_{i+1}(x_j) &= \ssc{s_i\alpha_j/\normm{\psi_j}_{L^1(\Omega)}}(\delta\gamma_i(x_j)-s_i\nabla_sR(\delta\gamma_i)(x_j)),\enskip j=1,2,\dots,N, \label{femupdate} \\
\delta\gamma_{i+1} &= \proja(\zeta_{i+1}). \notag
\end{align}
Note that the regularization parameter will depend quite heavily on the discretization of the mesh, i.e. for the same domain a good regularization parameter $\alpha$ will be much larger on a coarse mesh than on a fine mesh. This is quite inconvenient, and instead we can weigh the regularization parameter according to the mesh cells, by having $\alpha_j \equiv \alpha\beta_j\mu_j$. This leads to a discretization of a weighted $L^1$-norm penalty term:
\[
\alpha\int_{\Omega} f_\mu \absm{\dsr}\,dx \simeq \alpha\sum_j \beta_j\mu_j\absm{\dsr(x_j)},
\]
where $f_\mu : \Omega\to (0,1]$ is continuous and $f_\mu(x_j) = \mu_j$. For a triangulated mesh, the weights $\beta_j$ consists of the node area computed in 2D as 1/3 of the area of $\supp\psi_j$. This corresponds to splitting each cell's area evenly amongst the nodes, and it will not lead to instability on a regular mesh. This will make the choice of $\alpha$ almost independent of the mesh.
\begin{remark}
The corresponding algorithm with the FEM basis is the same as Algorithm \ref{alg1}, except that the update is applied via \eref{femupdate}.
\end{remark}
\section{Numerical Examples}\label{sec:NumericalResults}
In this section we illustrate, through several examples, the numerical algorithm implemented using the finite element library FEniCS \cite{logg2012a}. First we consider the full data case $\gamD = \gamN = \partial\Omega$ without and with prior information, and then we do the same for the partial data case. Finally, a brief comparison is made with another sparsity promoting method based on total variation.
For the following examples $\Omega$ is the unit disk in
$\rum{R}^2$. The regularization parameter $\alpha$ is chosen manually
by trial and error. The other parameters are $\sigma_0\equiv 1$, $M =
5$, $\tau = 10^{-5}$, $s_{\min} = 1$, $s_{\max} = 1000$, and the
stopping criteria is when the step size is below $s_{\text{stop}} =
10^{-3}$. $K = 10$ and the applied Neumann data will be of the form
$g_n^\textup{c}(\theta) \equiv \cos(n\theta)$ and
$g_n^\textup{s}(\theta) \equiv \sin(n\theta)$ for $n = 1,\dots,5$ and
$\theta$ being the angular variable. For the partial data an interval
$\Gamma = \gamN = \gamD = \{\theta\in(\theta_1,\theta_2)\}$ is
considered, and $g_n^\textup{c}$ and $g_n^\textup{s}$ are scaled and
translated such that they have $n$ periods in the interval.
When applying prior information, the coefficients $\mu_j$ are chosen as $10^{-2}$ where the support of $\dsr$ is assumed, and $1$ elsewhere. It should be noted that in order to get fast transitions for sharp edges when prior information is applied, a local mesh refinement is used during the iterations to refine the mesh where $\absm{\nabla\dsr}$ is large.
For the simulated Dirichlet data, the forward problem is computed on a very fine mesh, and afterwards interpolated onto a different much coarser mesh in order to avoid inverse crimes. White Gaussian noise has been added to the Dirichlet data $\{f_k\}_{k=1}^K$ on the discrete nodes on the boundary of the mesh. The standard deviation of the noise is chosen as $\epsilon \max_{k}\max_{x_j\in\gamD}\absm{f_k(x_j)}$ as in \cite{jin2012}, where the noise level is fixed as $\epsilon = 10^{-2}$ (corresponding to 1\% noise) unless otherwise stated.
\begin{figure}[h
\centering
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig2a}
\caption{{}} \label{fig:circphantom}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig2b}
\caption{{}} \label{fig:kitephantom}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig2c}
\caption{{}} \label{fig:hnlphantom}
\end{subfigure
\caption{\textbf{(a):} Circular piecewise constant inclusion. \textbf{(b):} Kite-shaped piecewise constant inclusion. \textbf{(c):} Multiple $C^2$ inclusions.}
\label{fig:phantoms}
\end{figure}
\Fref{fig:phantoms} shows the numerical phantoms: where one is a simple circular inclusion, another is the non-convex kite-shaped phantom. Finally, we also shortly investigate the case of multiple smoother inclusions.
\subsection{Full Boundary Data}
For $\gamD = \gamN = \po$ it is possible to get quite good reconstructions of both shape and contrast for the convex inclusions as seen in \fref{fig:recon1}, and for the case with multiple inclusions there is a reasonable separation of the inclusions.
\begin{figure}[h
\centering
\begin{subfigure}[h]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig3a}
\caption{$\alpha = 10^{-3}$}
\end{subfigure
\hfill
\begin{subfigure}[h]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig3b}
\caption{$\alpha = 5\cdot 10^{-4}$}
\end{subfigure
\hfill
\begin{subfigure}[h]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig3c}
\caption{$\alpha = 6.5\cdot 10^{-4}$}
\end{subfigure
\caption{Sparse reconstruction of the phantoms in \fref{fig:phantoms}.}
\label{fig:recon1}
\end{figure}
For the kite-shaped phantom we only get what seems like a convex approximation of the shape. It is seen in \cite{jin2012} that the algorithm is able to reconstruct some types of non-convex inclusions such as the hole in a ring-shaped phantom, however those inclusions are much larger which makes it easier to distinguish from similar convex inclusions.
We note that the method is very stable towards noise. In \fref{fig:recon2} it is shown how unreasonable amounts of noise only leads to small deformations in the shape of the reconstructed inclusion.
\begin{figure}[h
\centering
\begin{subfigure}[b]{.4\linewidth}
\includegraphics[width = \linewidth]{fig4a}
\end{subfigure
\hfill
\begin{subfigure}[b]{.3\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig4b}
\caption*{10\% noise}
\end{subfigure
\hfill
\begin{subfigure}[b]{.3\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig4c}
\caption*{50\% noise}
\end{subfigure
\caption{\textbf{Left:} Dirichlet data corresponding to $g = \cos(\theta)$ for the phantom in \fref{fig:circphantom}, with 10\% and 50\% noise level. \textbf{Middle:} reconstruction for 10\% noise level. \textbf{Right:} reconstruction for 50\% noise level.}
\label{fig:recon2}
\end{figure}
In order to investigate the use of prior information we consider the phantom in \fref{fig:circphantom}, and let $B(r)$ denote a ball centered at the correct inclusion and with radius $r$. Now we can investigate reconstructions with prior information assuming that the support of $\dsr$ is $\overline{B((1+\delta r)r^*)}$ for $r^*$ being the correct radius of the inclusion. \Fref{fig:r-dep} shows that underestimating the support of the inclusion $\delta r < 0$ is heavily enforced, and the contrast is vastly overestimated in the reconstruction as shown in \fref{fig:meanmax} (note that this can not be seen in \fref{fig:r-dep} as the color scale for the phantom is applied).
\begin{figure}[h
\centering
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig5a}
\caption*{No prior}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig5b}
\caption*{$\delta r = -0.25$}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig5c}
\caption*{$\delta r = -0.10$}
\end{subfigure
\\[5mm]
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig5d}
\caption*{$\delta r = 0$}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig5e}
\caption*{$\delta r = 0.10$}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig5f}
\caption*{$\delta r = 0.25$}
\end{subfigure
\caption{Sparse reconstruction of the phantom in \fref{fig:circphantom} for varying $\delta r$. The colorbar is truncated at $[1,6]$.}
\label{fig:r-dep}
\end{figure}
Interestingly, when overestimating the support, the contrast and support of the reconstructed inclusion does not suffer particularly. Intuitively, this corresponds to increasing $\delta r$ such that the assumed support of $\dsr$ contains the entire domain $\Omega$, which corresponds to the case with no prior information. For a subset $E\subseteq\Omega$ denote by $\sigma_E \equiv \absm{E}^{-1}\int_E \sigma dx$ the average of $\sigma$ on $E$, and denote by $\sigma_{\max} \equiv \max_{j}\absm{\sigma(x_j)}$ the maximum of $\sigma$ on the mesh nodes. Then \fref{fig:meanmax} gives a good indication of the aforementioned intuition, where around $\delta r = 0$ both $\sigma_B$ and $\sigma_{\max}$ levels off around the correct contrast of the inclusion (the red line) and stays there for $\delta r > 0$. It should be noted that even a 25\% overestimation of the support leads to a better contrast in the reconstruction than if no prior information was applied, as seen in \fref{fig:r-dep}.
Having an overestimation of the support for $\dsr$ also seems to be a reasonable assumption. Definitely there is the case of no prior information which means that $\supp \dsr$ is assumed to be $\Omega$. If the estimation comes from another method such as total variation regularization, then the support is typically slightly overestimated while the contrast suffers \cite{Borsic}. Thus we can use the overestimated support to get a good localisation and contrast reconstruction simultaneously.
\begin{figure}[h
\centering
\includegraphics[width = 0.6 \linewidth]{fig6}
\caption{Behaviour of sparsity reconstruction based on the phantom in \fref{fig:circphantom} for varying $\delta r$.}
\label{fig:meanmax}
\end{figure}
\Fref{fig:kitepriorrecon} shows how the reconstruction of the kite-shaped phantom can be vastly improved. Note that not only is $\supp \dsr$ better approximated, but the contrast is also highly improved. It is not surprising that we can achieve an almost perfect reconstruction if $\supp \dsr$ is exactly known, however it is a good benchmark to compare the cases for the overestimated support as it shows how well the method can possibly do.
\begin{figure}[h
\centering
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig7a}
\caption*{No prior}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig7b}
\caption*{10\% overestimated support}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 4cm 4cm 1cm 4cm, clip=true]{fig7c}
\caption*{Exact support}
\end{subfigure
\caption{Sparse reconstruction of the phantom in \fref{fig:kitephantom}.}
\label{fig:kitepriorrecon}
\end{figure}
\subsection{Partial Boundary Data}
For the partial data problem we choose $\Gamma = \gamD = \gamN = \{\theta\in(\theta_1,\theta_2)\}$ for $0\leq \theta_1 < \theta_2 \leq 2\pi$.
\begin{figure}[h
\centering
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.6cm 4cm 0cm 4cm, clip=true]{fig8a}
\caption*{Full data}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.6cm 4cm 0cm 4cm, clip=true]{fig8b}
\caption*{Top half}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.6cm 4cm 0cm 4cm, clip=true]{fig8c}
\caption*{Bottom half}
\end{subfigure
\caption{Sparse reconstruction of the phantom in \fref{fig:hnlphantom}. \textbf{Left:} $\Gamma=\po$. \textbf{Middle:} $(\theta_1,\theta_2) = (0,\pi)$. \textbf{Right:} $(\theta_1,\theta_2) = (\pi,2\pi)$.}
\label{fig:hnlpartial}
\end{figure}
In \fref{fig:hnlpartial} we observe that with data on the top half of the unit circle it is actually possible to get very good contrast and also reasonable localization of the two large inclusions. There is still a clear separation of the inclusions, while the small inclusion is not reconstructed at all. With data on the bottom half the small inclusion is reconstructed almost as well as with full boundary data, but the larger inclusions are only vaguely visible. This is the kind of behaviour that is expected from partial data EIT, and in practice it implies that we can only expect reasonable reconstruction close to where the measurements are taken.
\begin{figure}[h
\centering
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig9a}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig9b}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig9c}
\caption{{}}
\end{subfigure
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig9d}
\caption{{}}
\end{subfigure
\caption{Sparse reconstruction of the phantom in \fref{fig:circphantom}. \textbf{(a):} 50\% boundary data, no prior. \textbf{(b):} 50\% boundary data with 5\% overestimated support. \textbf{(c):} 25\% boundary data, no prior. \textbf{(d):} 25\% boundary data with 5\% overestimated support.}
\label{fig:circpartial}
\end{figure}
\begin{figure}[h
\centering
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig10a}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig10b}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig10c}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.25\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig10d}
\caption{{}}
\end{subfigure
\caption{Sparse reconstruction of the phantom in \fref{fig:kitephantom}. \textbf{(a):} 50\% boundary data, no prior. \textbf{(b):} 50\% boundary data with 10\% overestimated support. \textbf{(c):} 25\% boundary data, no prior. \textbf{(d):} 25\% boundary data with 10\% overestimated support.}
\label{fig:kitepartial}
\end{figure}
In \fref{fig:circpartial} and \fref{fig:kitepartial} panels (a) and (c) it is observed that as the length of $\Gamma$ becomes smaller, the reconstructed shape of the inclusion is rapidly deformed. By including prior information about the support of $\dsr$, it is possible to rectify the deformation of the shape, and get reconstructions with almost the correct shape but with a slightly worse reconstructed contrast compared to full boundary data reconstructions. This is observed for the ball and kite-shaped inclusions in \fref{fig:circpartial} and \fref{fig:kitepartial}.
\subsection{Comparison with Total Variation Regularization} \label{sec:comparison}
Another sparsity promoting method is total variation (TV) regularization, which promotes a sparse gradient in the solution. This can be achieved by minimizing the functional
\begin{equation}
\Psi_{\text{TV}}(\delta\gamma) \equiv \sum_{k=1}^K R_k(\delta\gamma) + P_{\text{TV}}(\delta\gamma), \enskip\delta\gamma\in\arum_0, \label{psieqTV}
\end{equation}
where the discrepancy terms $R_k$ remains the same as in \eqref{psieq}, but the penalty term is now given by
\begin{equation}
P_{\text{TV}}(\delta\gamma) \equiv \alpha\int_{\Omega} \sqrt{\absm{\nabla\delta\gamma}^2+b}\,dx.
\end{equation}
Here $b>0$ is a constant that implies that $P_{\text{TV}}$ is differentiable, but chosen small such that $P_{\text{TV}}$ approximates $\alpha\int_{\Omega} \absm{\nabla\delta\gamma}\,dx$.
For the numerical examples, the piecewise constant phantoms in \fref{fig:circphantom} and \fref{fig:kitephantom} are used, with the same noise level as in the previous sections. The value $b = 10^{-5}$ is used for the penalty term in all the examples.
It should be noted that the color scale in the following examples is not the same scale as for the phantoms, unlike the previous reconstructions. This is because the TV reconstructions have a significantly lower contrast, in particular for the partial data reconstructions, and would be visually difficult to distinguish from the background conductivity in the correct color scale.
\begin{figure}[h
\centering
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig11a}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig11b}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig11c}
\caption{{}}
\end{subfigure
\caption{TV reconstruction of the phantom in \fref{fig:circphantom}. \textbf{(a):} Full boundary data. \textbf{(b):} 50\% boundary data. \textbf{(c):} 25\% boundary data.}
\label{fig:circpartial2}
\end{figure}
\begin{figure}[h
\centering
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig12a}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig12b}
\caption{{}}
\end{subfigure
\hfill
\begin{subfigure}[b]{.33\linewidth}
\includegraphics[width = \linewidth, trim = 3.5cm 4cm 0cm 4cm, clip=true]{fig12c}
\caption{{}}
\end{subfigure
\caption{TV reconstruction of the phantom in \fref{fig:kitephantom}. \textbf{(a):} Full boundary data. \textbf{(b):} 50\% boundary data. \textbf{(c):} 25\% boundary data.}
\label{fig:kitepartial2}
\end{figure}
As seen from \fref{fig:circpartial2} and \fref{fig:kitepartial2} the support of the inclusion is slightly overestimated in the case of full boundary data, and for the partial data cases the support is slightly larger than the counterparts in \fref{fig:circpartial} and \fref{fig:kitepartial}. It is also noticed that the TV reconstructions have a much lower contrast than the $\ell_1$ sparsity reconstructions, and the contrast for the TV reconstructions is severely reduced when partial data is used. It is also observed that the same type of shape deformation occurs for both methods in case of partial data.
A typical feature of the TV regularization is piecewise constant reconstructions, however the reconstructions seen here have constant contrast levels with a smooth transition between them. There are several reason for this; and is due to the slight smoothing of the penalty term, but mostly because the discrepancy terms are not convex and may lead to local minima. The same kind of smooth transitions are also observed in TV-based methods for EIT in \cite{Borsic}.
\section{Conclusions}\label{sec:conclusions}
We have extended the algorithm developed in \cite{jin2012}, for sparse reconstruction in electrical impedance tomography, to the case of partial data. Furthermore, we have shown how a distributed regularization parameter can be applied to utilize spatial prior information. This lead to numerical results showing improved reconstructions for the support of the inclusions and the contrast simultaneously. The use of the distributed regularization parameter enables sharper edges in the reconstruction and vastly reduces the deformation of the inclusions in the partial data problem, even when the prior is overestimated.
The algorithm can be generalized for 3D reconstruction, under further assumptions on the boundary conditions $\{g_k\}_{k=1}^K$ and the amplitude of the perturbation $\delta\sigma$. This will be considered in a forthcoming paper \cite{GardeKnudsen2014}.
\section*{Funding}
This research is supported by Advanced Grant No.\ 291405 HD-Tomo from the European Research Council.
\bibliographystyle{gIPE}
|
\section{Introduction}
The curvature-square inflation originally proposed by Starobinsky
\cite{Starobinsky:1980te} occupies a unique position in inflationary
cosmology \cite{Guth:1980zm} because it only requires a single
additional term in the Einstein--Hilbert action instead of a new
scalar field---a functional degree of freedom. Despite its simplicity,
$R^2$ inflation is fully consistent with the state-of-the-art
cosmological observations of the cosmic microwave background (CMB)
by WMAP \cite{Hinshaw:2012aka}
and Planck \cite{Ade:2013uln}, which report the scalar spectral index
$n_s=0.963 \pm 0.014$ \cite{Ade:2013uln},
tensor-to-scalar ratio $r\lesssim 0.135$ \cite{Ade:2013uln,Audren:2014cea},
and the $f_{\rm NL}$ parameter measuring possible deviation from Gaussianity of curvature
perturbations being consistent with zero.
Recently, the BICEP2 experiment \cite{Ade:2014xna}
announced the detection of B-mode polarization of the
CMB on relatively low multipoles with its amplitude corresponding to
$r \sim 0.2$. This is an epoch-making discovery if it is confirmed to
be due to the primordial gravitational waves, but at the moment, the
possibility that the detected signal is entirely due to polarized dust
has been ruled out only at 2.2$\sigma$ level \cite{Ade:2014xna}. Furthermore, it has been
pointed out recently that the effect of foreground dust may have been
even larger, so that only an upper bound on $r$ can be obtained as reported in Ref.~\cite{seljak}.
In this sense, we had better not conclude in haste that
$R^2$ inflation, which predicts
$r \simeq 3\times 10^{-3}$, has been ruled out by the latest observations of
B-mode polarization.
Turning our attention to more theoretical aspects, this model is so simple
that it had not been investigated in the context of
modern high-energy theory including supersymmetry and supergravity,
which would be the only theories that allow us to use the usual perturbative quantum field theory
without a fine-tuning up to
scales relevant to inflation and necessary to embed it in the most promising candidate of the quantum
theory of gravity, the superstring theory.
It is only recently that $R^2$-type inflation was studied based on
supergravity
\cite{Kallosh:2013lkr,Farakos:2013cqa,Ferrara:2013kca,Hamaguchi:2014mza}\footnote{Another model of supergravity
extension of the Starobinsky model known as $F(R)$ supergravity
has also been proposed \cite{Ketov:2009wc} and its cosmological consequences have been studied
\cite{Watanabe:2013lwa}.
But its insufficiency as a supersymmetric theory has been pointed out recently \cite{Ferrara:2013wka,Ketov:2013dfa}.
See also \cite{Ellis:2013xoa} for
supersymmetric models that have a potential similar to the Starobinsky model.},
although its basic framework was already known
in the late 1980s using the old-minimal supergravity \cite{Cecotti:1987sa} or the new-minimal supergravity
\cite{Cecotti:1987qe}, which were obtained by the gauge fixing of the superconformal theory.
One of the most important messages of these studies is that higher-order corrections such as
the $R^4$ term will arise as the nonrenormalizable operators.
The effect of the $R^4$ term on the Starobinsky model can be seen easily when we go to the scalaron picture,
which is the dual theory of $f(R)$ theory that consists of a scalar field and the Einstein--Hilbert action.
In the scalaron picture, the Starobinsky model has a very flat potential at larger field values that is
suitable for inflation, but $R^4$ correction destroys the flatness of the potential.
Therefore, it must be strongly suppressed for the successful inflation \cite{Farakos:2013cqa,Ferrara:2013kca}.
In this paper, we investigate the quantitative constraint on the $R^4$ corrections to the Starobinsky model
in supergravity in light of the observational result and the possibility of its realization.
Even if the scalaron potential admits an inflationary solution, it is nontrivial whether it generates
the primordial perturbations consistent with the current observation.
As a result, we find severe constraints on the amplitude of the $R^4$ term.
Moreover, it is also nontrivial how severe tuning for the initial condition is required,
since the regions for successful inflation in the field space are drastically limited.
We find that the coupling of the $R^4$ term is very severely constrained if we start from the chaotic initial condition
\cite{Linde:1983gd}. However, we also find that if the scalaron potential vanishes at the larger field values,
depending on the sign of the coupling of the $R^4$ corrections, topological inflation \cite{Linde:1994hy,Sakai:1995nh,Ellis:1998gf} would be possible.
Therefore, the initial condition problem is solved in this case, and we should only focus on the observational constraint
for the embedding of the Starobinsky model in a supersymmetric theory.
Here, we take the old-minimal supergravity for concreteness, but
the same result is obtained in other supersymmetric extensions such as
the new-minimal supergravity, too, as far as
the form of the scalaron potential is concerned after all the other degrees of freedom have been stabilized.
Note that the observational constraint for the $R^4$ correction to the nonsupersymmetric Starobinsky model
has been studied in Ref.~\cite{Huang:2013hsb}. Our result is slightly different but consistent.
The paper is organized as follows.
In Sec.~\ref{sec2}, we derive the scalaron potential of the Starobinsky model in old-minimal supergravity
with $R^4$ correction. In Sec.~\ref{sec3}, we show the observational constraints to the $R^4$ correction
from the Planck results. We examine the initial condition problem from the chaotic initial condition and
show that the topological inflation likely takes place in Sec.~\ref{sec4}. Section~\ref{sec5} is devoted to the summary.
In the Appendix \ref{app}, we show the equivalence of the scalaron potential to the nonsupersymmetric Starobinsky model.
\section{Starobinsky model in old-minimal supergravity with higher-order corrections \label{sec2}}
We start from the old-minimal supergravity from superconformal gravity \cite{Cecotti:1987sa},
where the conformal symmetry is broken by the gauge fixing of a chiral compensator field\footnote{
In the new-minimal supergravity \cite{Cecotti:1987qe}, the conformal symmetry is broken by a real linear multiplet.}
following the
discussion of Ref.~\cite{Farakos:2013cqa}.
Here we introduce a chiral compensator superfield, $S_0$, for which the scaling weight is 1 and
chiral weight is 1/2.
Action in the superconformal theories can be categorized by the D-type and F-type Lagrangians.
The D-term Lagrangian is expressed as
\begin{equation}
{\cal L}_D=[V]_D=\int d^2 \Theta {\cal E}P[V]+{\rm h.c.},
\end{equation}
where $V$ is a real function of chiral fields for which the
scaling weight is 2 and chiral weight is 0,
${\cal E}$ is the chiral measure, and
$P$ is the chiral projector in conformal superspace.
The F-term Lagrangian is expressed as
\begin{equation}
{\cal L}_F+{\rm h.c.}=[W]_F+{\rm h.c.}=\int d^2 \Theta 2{\cal E} W +{\rm h.c.},
\end{equation}
where $W$ is a holomorphic function for which the scaling weight is 3 and chiral weight is 2.
The gravity part of the standard supergravity Lagrangian is obtained by the compensator chiral superfield as
\begin{equation}
{\cal L}=-3 [S_0{\bar S}_0]_D
\end{equation}
with the gauge fixing $S_0=1$\footnote{Here, we take the reduced Planck mass $M_{\rm pl}=1$.}.
Introducing a chiral multiplet
\begin{equation}
{\cal R}\equiv \frac{1}{2} S_0^{-1} P[{\bar S}_0],
\end{equation}
for which the scaling weight is 1 and chiral weight is 2/3, we obtain the $R^2$ correction to the standard supergravity Lagrangian,
\begin{equation}
{\cal L}=-3[S_0{\bar S}_0]_D+3 \lambda_1[{\cal R} {\bar {\cal R}}]_D. \label{orig}
\end{equation}
Note that after gauge fixing, the chiral projector is expressed as
\begin{equation}
P[V]=-\frac{1}{4}({\bar {\cal D}}{\bar {\cal D}}-8{\cal R}) V,
\end{equation}
where ${\cal D}$ is the covariant derivative,
and the ${\cal R}$ is the curvature chiral superfield.
As a result, the Lagrangian reads
\begin{equation}
{\cal L}=-3 \int d^2 \theta 2 {\cal E} \left\{{\cal R}+\frac{\lambda_1}{8}({\bar {\cal D}}{\bar {\cal D}}-8{\cal R})({\cal R} {\bar {\cal R}})\right\} +{\rm h.c.},
\end{equation}
which has been shown to be the supersymmetrization of the Starobinsky's $R^2$ model \cite{Ketov:2013dfa,Cecotti:1987sa}.
Now, let us consider higher-order corrections without fixing $S_0=1$ for the moment.
Since $[f({\cal R} {\bar {\cal R}})]_D$ corrections have been found not to lead the supersymmetrization of
$R^n$ corrections \cite{Ketov:2013dfa},
here we focus on the ghost-free correction that includes covariant derivatives of the curvature multiplet \cite{Khoury:2010gb},
\begin{equation}
\Delta {\cal L}=\xi \left[ \nabla^\alpha ({\cal R}/S_0) \nabla_\alpha ({\cal R}/S_0){\bar \nabla}_{\dot \alpha} ({\cal {\bar R}}/{\bar S}_0) {\bar \nabla}^{\dot \alpha} ({\cal {\bar R}}/{\bar S}_0) \right]_D, \label{corrrrr}
\end{equation}
where $\nabla$ represents the covariant derivative in the superconformal theory.
As we will see, the Lagrangian contains only first derivatives of the fields
and all the terms with the form of ${\cal L} \propto (\partial_\mu \phi)^2$ have the correct signs along
the inflationary trajectory. Thus,
this system does not suffer from the emergence of the ghost degrees of freedom. On the contrary,
terms expressed with a function $g$ as $[g(S_0^{-1}{\cal R}, {\bar S}_0^{-1}{\bar{\cal R}}, S_0^{-2}P({\bar {\cal R}}), {\bar S}_0^{-2} {\bar P}({\cal R})) S_0 {\bar S}_0]_D$ give ghost degrees of freedom \cite{Cecotti:1987sa,Ketov:2013dfa}, and hence
we do not consider them here.
Therefore, Eq.~\eqref{corrrrr} would be the lowest nonrenormalizable term that can give the consistent theory
in this framework.
Now, we examine the structure of this system by using the Lagrange multiplier method, or the scalaron picture.
By introducing a chiral superfield ${\cal A}$ with scaling weight 1 and chiral weight 2/3, and a chiral Lagrange
multiplier superfield $\Lambda$ with scaling weight 2 and chiral weight 4/3, the Lagrangian can be rewritten as
\begin{align}
{\cal L}&=-3[S_0{\bar S}_0]_D+3 \lambda_1[{\cal A}{\bar {{\cal A}}}]_D+\xi \left[ \nabla^\alpha ({\cal A}/S_0) \nabla_\alpha ({{\cal A}}/S_0) {\bar \nabla}_{\dot \alpha} ({\bar {\cal A}}/{\bar S}_0) {\bar \nabla}^{\dot \alpha} ( {\bar {\cal A}}/{\bar S}_0) \right]_D+3[\Lambda ({\cal A}-{\cal R})]_F+{\rm h.c.}
\end{align}
By integrating out the multiplier field, we have ${\cal A}={\cal R}$ and the original Lagrangian Eqs.~\eqref{orig} and \eqref{corrrrr}
are reproduced.
Noting that the identity $[\Lambda {\cal R}]_F+{\rm h.c.}=(1/2)[\Lambda S_0^{-1}{\bar S}_0+{\bar \Lambda} {\bar S}^{-1}_0 S_0]_D$ holds \cite{Cecotti:1987sa}, it can be further rewritten as
\begin{align}
{\cal L}=&-3[S_0{\bar S}_0-\lambda_1 {\cal A}{\bar {{\cal A}}} +(1/2)(\Lambda S_0^{-1}{\bar S}_0+{\bar \Lambda} {\bar S}^{-1}_0 S_0) ]_D \notag \\ &+\xi \left[ \nabla^\alpha ({\cal A}/S_0) \nabla_\alpha ({{\cal A}}/S_0) {\bar \nabla}_{\dot \alpha} ({\bar {\cal A}}/{\bar S}_0) {\bar \nabla}^{\dot \alpha} ( {\bar {\cal A}}/{\bar S}_0) \right]_D+3[\Lambda {\cal A}]_F+{\rm h.c.},
\end{align}
which will lead to the standard Poincar\'e supergravity with a chiral multiplet that has higher-order derivative coupling.
Defining new chiral multiplets
\begin{equation}
C\equiv \frac{\sqrt{\lambda_1}{\cal A}}{S_0}, \quad T=\frac{\Lambda}{2 S_0^2}+\frac{1}{2},
\end{equation}
we obtain the Lagrangian
\begin{equation}
{\cal L}=-3[S_0{\bar S}_0(T+{\bar T}-C{\bar C})]_D+\frac{\xi}{\lambda_1^2}\left[ \nabla^\alpha C \nabla_\alpha C {\bar \nabla}_{\dot \alpha} {\bar C}{\bar \nabla}^{\dot \alpha} {\bar C} \right]_D+\frac{6}{\sqrt{\lambda_1}}\left[S_0^3 C\left(T-\frac{1}{2}\right)\right]_F+{\rm h.c.}
\end{equation}
After gauge fixing $S_0=1$, the Lagrangian leads to,
\begin{align}
{\cal L}=&\int d^2 \Theta 2 {\cal E} \left[\frac{3}{8}\left({\bar {\cal D}} {\bar {\cal D}}-8{\cal R}\right)e^{-K/3}+W \right]+{\rm h.c.} \notag \\
&-\frac{\xi}{\lambda_1^2} \int d^2 \Theta 2 {\cal E} \left[ \frac{1}{8}\left({\bar {\cal D}} {\bar {\cal D}}-8{\cal R}\right) {\cal D}^\alpha C {\cal D}_\alpha C {\bar {\cal D}}^{\dot \alpha} {\bar C}{\bar {\cal D}}_{\dot \alpha} {\bar C}\right],
\end{align}
where
\begin{equation}
K\equiv -3\ln[T+{\bar T}-C{\bar C}], \quad W=\frac{6}{\sqrt{\lambda_1}}C\left(T-\frac{1}{2}\right).
\end{equation}
Expanding the component fields, integrating out the auxiliary fields in the gravity sector and
the F-term of the $T$ field, and performing an appropriate Weyl transformation, we have
\begin{align}
{\cal L}=\sqrt{-g}&\left[-\frac{R}{2}-K_{i{\bar j}}\partial_\mu z^i \partial^\mu z^{*j}-\frac{12}{\lambda_1}\frac{|C|^2}{T+T^*-|C|^2}\left(1-\frac{3(T+T^*-1)}{T+T^*-|C|^2}\right)\right. \notag \\
&+\frac{6}{\sqrt{\lambda_1}(T+T^*-|C|^2)^2}\left\{\left(|C|^2+T-\frac{1}{2} \right)F_C+{\rm h.c.}\right\}+\left(\frac{3}{(T+T^*-|C|^2)^2}-\frac{32\xi}{\lambda_1^2}\frac{\partial_\mu C \partial^\mu C^*}{T+T^*-|C|^2}\right)|F_C|^2 \notag \\
&\left.+\frac{16 \xi}{\lambda_1^2}\partial_\mu C \partial^\mu C \partial_\nu C^* \partial^\nu C^* +\frac{16\xi}{\lambda_1^2 (T+T^*-|C|^2)^2}|F_C|^4\right],
\end{align}
where $z^i=C,T$, $K_{i{\bar j}}\equiv \partial^2 K /\partial z^i \partial z^{*j}$ and $F_C$ is the
F-term of $C$.
Here we have used the same symbol for the superfield and its scalar component.
Since $K_{i{\bar j}}$ has the form
\begin{equation}
K_{T{\bar T}}=\frac{3}{(T+T^*-|C|^2)^2}, \quad K_{T{\bar C}}=\frac{-3C}{(T+T^*-|C|^2)^2},\quad K_{C{\bar C}}=\frac{3(T+T^*)}{(T+T^*-|C|^2)^2},
\end{equation}
the system does not have the ghost instability as long as $T+T^*>|C|^2$.
Note that the equation of motion for $C$ has only up to the second-order derivative, and hence
there arise no additional degrees of freedom.
$\partial {\cal L}/\partial F_C=0$ gives the condition that $F_C$ satisfies,
\begin{equation}
A+B F_C^*+2SF_C F_C^{*2}=0, \label{fcconst}
\end{equation}
where
\begin{align}
A&=\frac{6}{\sqrt{\lambda_1}(T+T^*-|C|^2)^2}\left(|C|^2+T-\frac{1}{2}\right), \\
B&=\left(\frac{3}{(T+T^*-|C|^2)^2}-\frac{32\xi}{\lambda_1^2}\frac{\partial_\mu C \partial^\mu C^*}{T+T^*-|C|^2}\right), \\
S&=\frac{16\xi}{\lambda_1^2 (T+T^*-|C|^2)^2}.
\end{align}
Then, $|F_C|^2$ satisfies the equation
\begin{equation}
\alpha=(1+\beta|F_C|^2)^2|F_C|^2,
\end{equation}
with
\begin{equation}\alpha=\frac{|A|^2}{B^2}, \quad \beta=\frac{2S}{B}.
\end{equation}
Here $\alpha$ is always positive, and
assuming that $\partial_\mu C=0$, the sign of $\beta$ is determined by $\xi$.
In the case $\beta>0$,
Eq.~\eqref{fcconst} has only one real and positive solution,
\begin{equation}
|F_C|^2=\frac{2}{3 \beta}(\cosh m -1), \label{fterm}
\end{equation}
where
\begin{equation}
m=\frac{1}{3}\cosh^{-1}\left(\frac{27}{2}\alpha \beta+1\right).
\end{equation}
Note that $1+(27/2)\alpha \beta>1$ is always satisfied in this case.
On the other hand, in the case $\beta<0$, the situation is relatively complicated.
If $0<\alpha<-(4/27)\beta^{-1}$, or $-1<1+(27/2)\alpha \beta<1$, Eq.~\eqref{fcconst} has three real and positive solutions,
\begin{equation}
|F_C|^2=\left\{
\begin{array}{l}
\dfrac{2}{3 \beta}(\cos {\tilde m} -1) \\ \\
\dfrac{2}{3 \beta}\left(\cos \left({\tilde m}+\dfrac{2\pi}{3}\right) -1\right) \\ \\
\dfrac{2}{3 \beta}\left(\cos \left({\tilde m}-\dfrac{2\pi}{3}\right) -1\right)
\end{array}\right.,
\end{equation}
where
\begin{equation}
{\tilde m}=\frac{1}{3}\cos^{-1}\left(\frac{27}{2}\alpha \beta+1\right).
\end{equation}
If $\alpha>(4/27)\beta^{-1}$ it has again only one real and positive solution,
\begin{equation}
|F_C|^2=\frac{1}{3 \beta} (-2+Z^{1/3}+Z^{-1/3}),
\end{equation}
with
\begin{equation}
Z\equiv\frac{2+27 \alpha \beta+\sqrt{27 \alpha \beta(4+27\alpha \beta)}}{2}.
\end{equation}
Let us study the resultant Lagrangian.
The full Lagrangian in which all the auxiliary fields are integrated out is
\begin{align}
{\cal L}=\sqrt{-g}&\left[-\frac{R}{2}-K_{i{\bar j}}\partial_\mu z^i \partial^\mu z^{*j}+\frac{16 \xi}{\lambda_1^2}\partial_\mu C \partial^\mu C \partial_\nu C^* \partial^\nu C^* \right. \notag \\
&\left.-\frac{12}{\lambda_1}\frac{|C|^2}{T+T^*-|C|^2}\left(1-\frac{3(T+T^*-1)}{T+T^*-|C|^2}\right)-B|F_C|^2-3S|F_C|^4\right], \label{fulllag}
\end{align}
independent of the value of $\beta$ with F terms given above.
Taking $C=0$\footnote{For small $\xi$, the Lagrangian \eqref{fulllag}
reveals a tachyonic instability for $C$ \cite{Farakos:2013cqa}. However, by introducing
the $[({\cal R}{\bar {\cal R}})^2/(S_0{\bar S}_0)]_D \rightarrow [(C{\bar C})^2]_D$ term, $C$ can acquire a positive mass squared
and the tachyonic instability problem can be solved \cite{Kallosh:2013lkr,Farakos:2013cqa}.
Here we assume implicitly such an extra term.
Note that such a term does not change the Lagrangian for $T$ in the $C=0$ direction. We discuss it
in more detail in Appendix~\ref{app2}. },
the Lagrangian is now of the form
\begin{align}
{\cal L}=\sqrt{-g}&\left[-\frac{R}{2}-\frac{3}{(T+T^*)^2}\partial_\mu T \partial^\mu T^*-B|F_C|^2-3S|F_C|^4\right].
\end{align}
Let us define
\begin{equation}
T=\frac{1}{2}e^{\sqrt{2/3}\phi }+ ib,
\end{equation}
to canonicalize the real part of the $T$ field.
Noting that we now have
\begin{align}
B&=\frac{3}{(T+T^*)^2}=3 e^{-2 \sqrt{2/3}\phi}, &S&=\frac{16 \xi}{\lambda_1^2(T+T^*)^2}=\frac{16 \xi}{\lambda_1^2} e^{-2 \sqrt{2/3}\phi}, \notag \\
\alpha&=\frac{4}{\lambda_1}\left|T-\frac{1}{2}\right|^2=\frac{(e^{\sqrt{2/3}\phi}-1)^2+4b^2}{\lambda_1}, &\beta&=\frac{32\xi}{3\lambda_1^2},
\end{align}
the Lagrangian becomes
\begin{align}
{\cal L}=\sqrt{-g}&\left[-\frac{R}{2}-\frac{1}{2}\partial_\mu \phi \partial^\mu \phi-3e^{-2\sqrt{2/3}\phi}\partial_\mu b \partial^\mu b -V\right]
\end{align}
where
\begin{equation}
V(\phi)=\frac{3\lambda_1^2}{16 \xi} e^{-2\sqrt{2/3}\phi} X(X-1) \label{potential}
\end{equation}
with
\begin{equation}
|F_C|^2=\frac{2}{3 \beta}(X-1).
\end{equation}
The expression of $X$ is different depending on the values of $\xi$ and $\phi$ as
\begin{equation}
X= \left\{
\begin{array}{ll}
\cosh m & \text{for} \quad \xi>0, \\
\left(
\begin{array}{l}
\cos {\tilde m} \\
\cos \left({\tilde m}+2\pi/3\right) \\
\cos \left({\tilde m}-2\pi/3\right)
\end{array} \right. & \text{for} \quad \xi<0 \quad \text{and} \quad \phi<\sqrt{\dfrac{3}{2}}\log \left[ 1+\dfrac{1}{6 \sqrt{2|s|}}\right] \equiv \phi_c \\
(Z^{1/3}+Z^{-1/3})/2 & \text{for} \quad \xi<0 \quad \text{and} \quad \phi >\phi_c
\end{array} \right. , \label{XX}
\end{equation}
where
\begin{equation}
s\equiv \frac{\xi}{\lambda_1^3}.
\end{equation}
Here, the $X=\cos({\tilde m}+2\pi/3)$ branch for $\xi<0$ smoothly connects to the solution for $\phi>\phi_c$.
$m$ and ${\tilde m}$ are expressed by $\phi$ and $b$ as
\begin{align}
m&=\frac{1}{3}\cosh^{-1}\left[144s((e^{\sqrt{2/3}\phi}-1)^2+4b^2)+1\right], \label{m1} \\
{\tilde m}&=\frac{1}{3}\cos^{-1}\left[144 s((e^{\sqrt{2/3}\phi}-1)^2+4b^2)+1\right], \label{m2}
\end{align}
and the condition $\alpha <-(4/27)\beta^{-1}$ yields $\phi<\phi_c$.
One may wonder which branch to take for $s<0$.
We find that the branch $X= \cos {\tilde m}$ has the potential minimum $V=0$ at $\phi=0$ for $b=0$ and
approaches the pure Starobinsky model for the $s \rightarrow 0$ limit,
whereas other branches as well as the solution $\phi>\phi_c$
have no potential minimum, and
the potential takes a negative value at $\phi \rightarrow -\infty$.
Therefore, we take the branch $X= \cos {\tilde m}$ as the supresymmetrized Starobinsky model
with an $R^4$ correction for $s<0$ and $\phi<\phi_c$.
Since other branches do not have well-defined vacua, hereafter we do not consider them.
In Appendix~\ref{app}, we show that the resultant potential is the equivalent to the
Starobinsky model with a $R^4$ correction in the nonsupersymmetric case, which strongly suggests that the model is
its supersymmetrized one.
Figure \ref{fig:1} shows the parameter dependence of the potential shape for $b=0$.
For $s>0$, the potential has a maximum and approaches to $V=0$ at $\phi=0$ and $\phi \rightarrow \infty$.
On the other hand, for $s<0$, it is a continuously increasing function with respect
to $\phi$ and undefined for $\phi>\phi_c$.
In both cases, the flatness of the potential appears to be violated for $|s|>10^{-7}$,
making it difficult for inflation to take place.
We will see how inflation can take place and how the correction is constrained observationally
in the next section.
\begin{figure}[t]
\center
\includegraphics[width = 0.8\textwidth]{potentialplot2.pdf}
\caption{The potential with higher-order correction for the Starobinsky model with various amplitudes of the corrections is shown.
Here we take $\lambda_1=8 \times 10^{10}(M_{\rm pl}^{-2})$.}
\label{fig:1}
\end{figure}
Here we comment on the $b$ field.
Since the imaginary part $b$ receives a positive mass squared for $\phi>0$ larger than $H^2$ along the
inflationary trajectory,
we can safely take $b=0$ and we have only to focus on the dynamics of $\phi$ field.
In Appendix~\ref{app2}, we examine it in detail.
\section{Observational constraints \label{sec3}}
Now, let us consider inflation driven by the $\phi$ field and study the observational constraints on the model parameters.
Inflation takes place when the slow-roll conditions,
\begin{equation}
\epsilon \equiv \frac{1}{2}\left(\frac{\partial V/\partial \phi}{V}\right)^2 \ll 1, \quad |\eta| \equiv \frac{1}{V}\left|\frac{\partial^2 V}{\partial \phi^2}\right| \ll1,
\end{equation}
are satisfied.
We can easily find that, even for relatively large $|s| \gg 1$, there are field spaces in which
slow-roll conditions $\eta \ll 1, \epsilon \ll 1 $ are simultaneously satisfied.
Therefore, slow-roll inflation can take place naturally in a sense apart from the observational consequences
and initial condition problem.
We will turn to the latter in the next section.
During inflation, the system obeys the slow-roll equations,
\begin{equation}
3H^2=V(\phi), \quad \frac{d \phi}{dN}=\frac{\partial V/\partial \phi}{V},
\end{equation}
where
$N$ is defined as $dN=-Hdt$, and inflation ends when
$\epsilon$ reaches unity
at $\phi=\phi_{\rm f}$.
Cosmological perturbations are generated during inflation. They are quantified in terms of
the amplitude of the scalar fluctuations $A_s$, the scalar spectral index $n_s$ and the tensor-to-scalar ratio $r$,
\begin{align}
A_s&= \frac{H^2}{8\pi \epsilon}, \label{asas} \\
n_s&=1-6 \epsilon+2 \eta, \\
r&=16 \epsilon,
\end{align}
which are evaluated at the $\phi$ field value when the relevant scale leaves the horizon during inflation.
If there are no additional sources of cosmological perturbations,
they are directly compared to the Planck and other cosmological observations.
We adopt the $\phi$ field value at the number of $e$-folds at $N_*\simeq 55$ before the end of inflation when the pivot scale
$k_*=0.05 \rm{Mpc}^{-1}$ leaves the horizon\footnote{See Ref.~\cite{Audren:2014cea} for the discussion
of the pivot scale in light of the BICEP2 result.}.
The model parameters, $\lambda_1$ and $\xi$, are constrained by the observations \cite{Ade:2013uln, Audren:2014cea},
\begin{equation}
A_s^{\rm obs} =(2.18 \pm 0.05)\times 10^{-9}, \quad n_s^{\rm obs} =0.963 \pm 0.007, \quad r<0.135. \label{planckres}
\end{equation}
For $s=0$, the scalaron potential\footnote{The dynamics of scalaron
oscillation with this potential is investigated in Ref.~\cite{Takeda:2014qma}
.} is given by
\begin{equation}
V(\phi)=\frac{3}{\lambda_1}(1-\exp[-\sqrt{3/2}\phi])^2,
\end{equation}
which yields $3 H^2 \simeq 3/\lambda_1$ during inflation,
and solving the slow-roll equations analytically, we obtain
\begin{equation}
\epsilon \simeq \frac{3}{4N_*^2}.
\end{equation}
Then, comparing Eqs.~\eqref{asas} and \eqref{planckres}, we find
\begin{equation}
\lambda_1 =\frac{N_*^2}{6 \pi A_s^{\rm obs}} \simeq 7 \times 10^{10}.
\end{equation}
Since the slow-roll dynamics of $\phi$ cannot be solved analytically when the the higher-order
correction exists, we have performed numerical calculation of the inflationary dynamics and evaluated the
primordial perturbations with various values of $\lambda_1$ and $s$.
Parameter regions that are favored by Planck are shown in Fig. \ref{fig:2} in the cases $s<0$ and $s>0$.
In both cases, for $|s|>10^{-7}$, the higher-order corrections are no longer negligible
for the inflaton dynamics, and hence the predictions start to deviate from the
pure Starobinsky model's; for $A_s=A_s^{\rm obs}$, $\lambda_1\simeq 7 \times 10^{10} $ is required,
and for these parameter values, the scalar spectral index and the tensor-to-scalar ratios are
predicted as $n_s \simeq 0.963$ and $r\simeq 3\times 10^{-3}$.
As a result, the value of $s$ is constrained as
\begin{equation}
-5.5 \times 10^{-8} < s<9.1\times 10^{-8},
\end{equation}
by the Planck observation at the $2 \sigma$ confidence level.
Therefore we conclude that the amplitude of the parameter $s$ must be smaller than at least
$10^{-7}$ to explain the current Universe in the context of the supergravity
Starobinsky model.
This means that some symmetries or mechanisms to reduce the higher-order corrections
to the Starobinsky model up to this level are necessary to derive it from the physics in the higher energy scales.
Here, we explain the behaviors of the parameter dependence of the observables.
In the case with $s<0$, larger $|s|$ leads to larger values of $\epsilon$ during inflation.
Therefore, larger potential energy or smaller $\lambda_1$ is required to generate the correct amplitude
of the scalar perturbations $A_s$, which leads to relatively large values of the tensor-to-scalar ratio $r$.
At the same time, the slow-roll parameter $\eta$ becomes a larger or even positive value, which leads to a
larger value of the scalar spectral index $n_s$.
On the other hand, in the case with $s>0$, larger $s$ leads to smaller values of $\epsilon$ during inflation.
This leads to larger values of $\lambda_1$ to generate the correct amplitude of $A_s$, which
also means smaller values of $r$. Simultaneously, for larger $s$, $\eta$ becomes larger,
which leads to a smaller value of $n_s$.
Note that the slow-roll parameters are independent of $\lambda_1$, and hence the observables
$n_s$ and $r$ are $\lambda_1$ independent.
\begin{figure}[t]
\centering{
\includegraphics[width = 0.45\textwidth]{pararegionplot4.pdf}
\includegraphics[width = 0.45\textwidth]{pararegionplot3.pdf}
}
\caption{The cosmological perturbations generated from the supersymmetric Starobinsky model
with higher-order corrections with respect to the model parameters are shown;
({\it left}) in the case with $s<0$ and ({\it right}) in the case with $s>0$.
Curved (red) lines represent the contour lines of the amplitude of the scalar fluctuation $A_s$.
The thick (light) shaded (blue) regions are the regions in which the values of the scalar spectral index are within the
Planck $1 \sigma (2\sigma)$ constraints. Horizontal (green) lines are the contour lines for the typical values of
the tensor-to-scalar ratio $r$. }
\label{fig:2}
\end{figure}
\section{Initial condition for inflation \label{sec4}}
Now, let us consider the initial condition problem, which is
strikingly different depending on the sign of $s$.
First, for $s<0$, as we have seen above, the field range that can evolve
into the proper vacuum after inflation is limited to $\phi < \phi_c$.
For a sufficient amount of inflation, slow-roll inflation should start
at $\phi \gtrsim 6$. Hence, severe fine-tuning of the initial condition
at, say, the Planckian epoch is necessary for both $\phi$ and $\dot\phi$.
If $\dot\phi$ has a Planckian value $\dot\phi \sim 1$
initially, the scalar field amplitude varies $\Delta\phi \sim 10$ before
the slow-roll inflation phase sets in.
Therefore, $\phi_c$ must be larger than $15-16$ for this initial velocity,
which turns to the constraint on $s$ as $|s|<3.2 \times 10^{-13}$.
For the larger amplitude of $s$,
the initial velocity must be suppressed accordingly.
On the other hand, for $s>0$, there is no restriction in the field range
of $\phi$, and the potential has a local maximum at $\phi \equiv \phi_t
\simeq -0.93\log_{10}(3.5s)$.
Hence, if the universe starts with a chaotic initial condition, some
domain falls into $\phi=0$, and others run away to infinity.
Note that since the $C$ and $b$ fields are stabilized at the origin for
any field values of $\phi$
we can take $C=b=0$ in all the domains.
Between these domains with different fates exists a region trapped to the potential maximum at $\phi=\phi_t$,
namely, a domain wall.
As Vilenkin and Linde have pointed out \cite{Linde:1994hy}, inflation can naturally take place
inside the domain wall, where the large energy density distributes relatively homogeneously,
if its thickness is larger than the local Hubble radius.
This is so-called the ``topological inflation.''
Thus, in the present case, it may be possible for the topological
inflation to take place.
Let us study this possibility in detail.
The condition for the realization of topological inflation is numerically studied in Ref.~\cite{Sakai:1995nh}
in the case with a potential $V=\kappa(\phi^2-v^2)^2$, and it was concluded that
a domain wall triggers inflation if $v$ is larger than the critical value $v_c \equiv 1.7$ regardless of
the value of $\kappa$. Since the potential we are studying is different from the double-well type,
the conclusion in Ref.~\cite{Sakai:1995nh} cannot be applied directly.
However, it is plausible that inflation can take place from the domain wall in the following reasons.
The thickness of the domain wall can be evaluated as $\delta = |V''(\phi_t)|^{-1/2}$.
Since numerically we find that
\begin{equation}
V''(\phi_{\rm t})\simeq -2.1 \times 10 \frac{s^{1/3}}{\lambda_1}
\end{equation}
and the Hubble parameter is evaluated as $H= (V(\phi_{\rm t})/3)^{1/2} \simeq 1/\lambda_1^{1/2}$,
we have the relation
\begin{equation}
H \delta \simeq 0.22 s^{-1/6}.
\end{equation}
Since the condition for the critical ratio between the wall thickness to the Hubble length given in Ref.~\cite{Sakai:1995nh} is
$H \delta =0.48$, the one in our case for $s<10^{-7}$ is much larger than the critical value.
Figure~\ref{fig:5} shows the shape of potential and its second derivative both in the case with our potential
with $s=10^{-7}$ and the double-well potential with $v\simeq 6$, both of which have the same potential maximum.
We can see that our potential is flatter than the double-well potential that can trigger the topological inflation.
Therefore, topological inflation will naturally take place in our potential
satisfying the observational constraint $s < 9.1 \times 10^{-8}$.
\begin{figure}[t]
\centering{
\includegraphics[width = 0.45\textwidth]{potentialdoublewell.pdf}
\includegraphics[width = 0.45\textwidth]{potentialderivl.pdf}
}
\caption{The potential ({\it left}) and its second derivative with respect to $\phi$ ({\it right}) of both the Starobinsky model
with higher order corrections $s=10^{-7}$ and the double well potential are shown.
The double potential has the maximum at $\phi=6$ and the amplitude is the same to the Starobinsky model.
Around the potential maximum, the Starobinsky model is flatter than the double-well potential. }
\label{fig:5}
\end{figure}
\section{summary \label{sec5}}
Starobinsky's $R^2$ inflation is one of the most attractive inflation models in light of the
Planck result. However,
the mechanism to induce the correct $R^2$ term that explains the observational result is not known.
Therefore, it would be a good direction to embed it in a supersymmetric theory
because it is one of the most promising physics beyond the Standard Model
and would be the key to the quantum theory of gravity.
On the other hand, once we consider the supersymmetric theory of the $R^2$ model,
higher-order terms cannot be forbidden by symmetry.
In this paper, we have studied the Starobinsky model in the old-minimal supergravity
with an $R^4$ correction that is free from ghost degrees of freedom.
After confirming that fields other than the scalaron field are stabilized appropriately, we focused on the dynamics of the scalaron sector.
Since the $R^4$ correction easily violates the flatness of the inflaton potential
in the scalaron picture, it should be strongly constrained.
We find that the constraint on the $R^4$ term is not so strong just for
the accelerating expansion of the Universe, but in order to generate the
spectral index of primordial scalar perturbation that is consistent with
Planck result, it is strongly constrained.
It is found that in terms of dimensionless coupling constant $s \equiv \xi/\lambda_1^3$
it is constrained as
\begin{equation}
-5.5 \times 10^{-8} <s < 9.1 \times 10^{-8}.
\end{equation}
On the initial condition, we also find the difficulties in the realization of inflation
when there is an $R^4$ correction.
From the chaotic initial condition in which the Universe starts from the Planck scale,
the $R^4$ term must be very severely constrained for $s<0$,
where the scalaron potential jumps up at the field value larger than the value
at which the $R^4$ correction becomes dominant.
On the other hand, in the case of $s>0$, the shape of the scalaron potential is
hilltop type, and domain walls are generated somewhere in the Universe regardless of the initial condition.
We find that the domain wall is thick enough for the topological inflation
for $s<10^{-7}$, and hence we do not suffer from the initial condition problem
in this case.
In summary, for the reasonable initial conditions, the $R^4$ correction is
constrained as
\begin{equation}
-3.2 \times 10^{-13}\ll s < 9.1 \times 10^{-8}
\end{equation}
for the realization of inflation that leads to the present Universe.
This would be an important constraint for the embedding or inducing
the Starobinsky model of inflation from the high-energy theory.
\begin{acknowledgments}
K.K. is grateful to K.~Ohashi, R.~Rattazzi, and A.~Westphal for useful comments.
The work of K.K. is supported by a JSPS postdoctoral fellowship for
research abroad. The work of J.Y. is supported by JSPS
Grant-in-Aid for Scientific Research, Grant No.~23340058.
\end{acknowledgments}
|
\section{Introduction}
\subsection{\hspace{0.1pt} }
Let $\mathfrak g$ be a finite-dimensional semisimple Lie algebra over $\mathbb{C}$.
The quantized enveloping algebra $U_q(\mathfrak g)$ was introduced in \cite{Dr-quantum,J}
by deforming the relations of the enveloping algebra $U(\mathfrak g)$.
If the parameter $q$ is not a root of unity, then $U_q(\mathfrak g)$ (defined as in \cite{DP-quantum}) has similar properties to $U(\mathfrak g)$.
But if $q$ is a root of unity, then $U_q(\mathfrak g)$ differs significantly from $U(\mathfrak g)$,
For instance
there is a Hopf pairing between $U_q^{+}(\mathfrak g)$ (the subalgebra generated by the $E_i$'s)
and $U_q^{-}(\mathfrak g)$ (the subalgebra generated by the $F_i$'s) but is degenerate
and its radical is the ideal generated by powers of roots vectors $E_\alpha^N$,
respectively $F_\alpha^N$, $\alpha\in\Delta_+$. Here $E_\alpha$, $F_\alpha$, $\alpha\in\Delta_+$,
are obtained from $E_i$, $F_i$ by applying Lusztig isomorphisms $T_i\in \Aut U_q(\mathfrak g)$ \cite[Section 9]{DP-quantum}.
The quotient of $U_q(\mathfrak g)$ by the ideal generated by $E_\alpha^N$, $K_\alpha^N$, $F_\alpha^N$ is a finite-dimensional Hopf algebra
known as the \emph{small quantum group} or \emph{Frobenius-Lusztig kernel} $\mathfrak{u}_q(\mathfrak g)$.
Now the induced Hopf pairing between $\mathfrak{u}_q^{+}(\mathfrak g)$ and $\mathfrak{u}_q^{-}(\mathfrak g)$ is non-degenerate, so that $\mathfrak{u}_q^{+}(\mathfrak g)$ is a Nichols algebra
\cite{AS Pointed HA}.
The kernel of the natural projection $U_q(\mathfrak g) \to \mathfrak{u}_q(\mathfrak g)$ (in the Hopf algebra sense) is the central Hopf subalgebra $Z_0$ generated
by $E_\alpha^N$, $K_\alpha^N$, $F_\alpha^N$; $Z_0$ is the algebra of functions of a Poisson group \cite[Chapter 19]{DP-quantum}
that plays roles in the representation theory of $U_q(\mathfrak g)$ \cite{DP-quantum, DPRR} and in the classification of finite-dimensional pointed Hopf algebras
\cite{AS Class}.
\subsection{\hspace{0.1pt} }
A \emph{braided vector space} is a pair $(V,c)$, where $V$ is a vector space and $c\in\Aut(V\otimes V)$ satisfies
$(c\otimes \id )(\id \otimes c)(c\otimes \id )=(\id \otimes c)(c\otimes \id )(\id \otimes c)$. The Nichols algebra of a braided vector space $(V,c)$
is a graded braided Hopf algebra $\mathcal{B}(V) = T(V)/\mathcal{J}(V) = \oplus_{n\ge 0}\mathcal{B}^n(V)$ with the property
that all primitive elements are in degree 1.
The study of Nichols algebras is crucial in the classification program of Hopf algebras \cite{AS Pointed HA}.
If there are a basis $(v_i)_{1\le i\le\theta}$ of $V$ and a matrix $(q_{ij})\in (\mathbb{C}^{\times})^{\theta\times \theta}$
such that $c(v_i\otimes v_j)=q_{ij} \, v_j\otimes v_i$ for all $1\le i,j\le\theta$, then $(V,c)$ is of \emph{diagonal type}.
Finite-dimensional Nichols algebras of (braided vector spaces of) diagonal type are classified in \cite{H-classif RS};
the defining relations of them are described in \cite{A-convex,A-presentation}.
A pre-Nichols algebra\footnote{This terminology is due to Akira Masuoka.} of a braided vector space $(V,c)$ is any graded braided Hopf algebra intermediate between $T(V)$ and $\mathcal{B}(V)$,
that is any braided Hopf algebra of the form $T(V)/\mathcal{I}$ where $\mathcal{I} \subseteq \mathcal{J}(V)$ is a homogeneous Hopf ideal.
Pre-Nichols algebras form a partially ordered set with order given by projection,
with $T(V)$ the minimal and $\cB(V)$ the maximal points.
Pre-Nichols algebras appear naturally in the computation of the deformations or liftings of \cite{Masuoka, AAGMV}.
Let $V$ be a braided vector space of diagonal type with finite-dimensional $\mathcal{B}(V)$.
In this paper we define and investigate the \emph{distinguished} pre-Nichols algebra $\widetilde{\mathcal{B}}(V)$.
Actually they are already present without name in \cite[Proposition 3.3]{A-presentation}.
The distinguished pre-Nichols algebra $\widetilde{\mathcal{B}}(V)$ can be realized in the category of Yetter-Drinfeld modules
over $ \mathbf{k}\mathbb{Z}^{\theta}$, where $\theta = \dim V$; let $U(V)$ be the quantum double of the bosonization
$\widetilde{\mathcal{B}}(V)\# \mathbf{k}\mathbb{Z}^{\theta}$. Similarly, let $\mathfrak{u}(V)$ be the quantum double of the bosonization
$\mathcal{B}(V)\# \mathbf{k}\mathbb{Z}^{\theta}$.
Here are some properties of $\widetilde{\mathcal{B}}(V)$, justifying the adjective distinguished (to our understanding):
\subsubsection{} Let $i\in\I_{\theta}$ and $\rho_i: V \mapsto V'$ the corresponding reflection in the Weyl groupoid of $(V,c)$.
The Lusztig isomorphism $T_i:\mathfrak{u}(V) \to \mathfrak{u}(V')$ can be lifted to a Lusztig isomorphism $T_i: U(V)\to U(V')$ \cite[Proposition 3.26]{A-presentation}.
We conjecture that $\widetilde{\mathcal{B}}(V)$ is minimal among the pre-Nichols algebras admitting Lusztig isomorphisms.
\subsubsection{} If $\cB (V) = \mathfrak{u}_q^{+}(\mathfrak g)$, then $\widetilde{\mathcal{B}}(V) = U_q^{+}(\mathfrak g)$.
\subsubsection{} By a general result of \cite{Kh}, every pre-Nichols algebra has a restricted (that is, with heights) PBW basis.
We prove the existence of a PBW basis with the same generators (root vectors), but different heights,
as $\cB(V)$, obtained by applying several $T_i$'s, see Theorem \ref{thm: PBW bases preNichols}.
\subsubsection{} The pointed Hopf algebra $U(V)$ is Noetherian of finite Gelfand-Kiri\-llov dimension, see Theorems \ref{thm:noetheriano} and \ref{thm:GKdim finita}.
We conjecture\footnote{jointly with N. A.} that $\widetilde{\mathcal{B}}(V)$ is minimal among the pre-Nichols algebras with these properties.
Observe that $\widetilde{\mathcal{B}}(V)$ is not a domain in general.
\subsubsection{} The powers of root vectors that are non-zero in the $\widetilde{\mathcal{B}}(V)$ but zero in $\cB(V)$
generate a subalgebra $Z^+(V)$ on $\widetilde{\mathcal{B}}(V)$, that coincides with the intersection of the kernels of the
skew-derivations associated to the coproduct of $U^+(V)$, see Theorem \ref{thm:Z+ es subalg hopf U+}. Correspondingly there is a normal Hopf subalgebra $Z(V)$ of $U(V)$,
see Theorem \ref{thm:Z es subalg hopf U};
$U(V)$ is a finite free $Z(V)$-module.
\subsection{\hspace{0.1pt} }The organization of the article is the following. In Section \ref{section:preliminares} we recall notions and
basic properties of generalized root systems and Lusztig isomorphisms of quantum doubles of Nichols algebras.
In Section \ref{section:Distinguished pre-Nichols algebras} we study the distinguished pre-Nichols algebras $\widetilde{\mathcal{B}}(V)$.
First we prove the existence of Lusztig isomorphisms for the quantum double $U(V)$ and PBW bases
whose PBW generators are obtained by applying Lusztig isomorphisms. Then we obtain an algebra filtration on $\widetilde{\mathcal{B}}(V)$ and $U(V)$ such that the associated graded algebra is a quantum polynomial algebra, so $\widetilde{\mathcal{B}}(V)$ and $U(V)$ are Noetherian of finite Gelfand-Kirillov dimension.
In Section \ref{section:subalgebra Z} we study the subalgebra $Z(V)$ of $U(V)$ generated by powers of root vectors.
First we show that each element of $Z(V)$ commutes up to scalars with each homogeneous element of $U(V)$, so $U(V)$ is a free $Z(V)$-module.
Next we obtain a formula relating the coproduct on $\widetilde{\mathcal{B}}(V)$ with Lusztig isomorphisms, close to \cite[Theorem 4.2]{HS-coideal subalg}.
We present a recursive formula for the coproduct of powers of root vectors in $\widetilde{\mathcal{B}}(V)$, with a view towards
the computation of the liftings of $\cB(V)$, as proposed in \cite{AAGMV}.
Finally, we compute the coproduct of $Z(V)$ for braidings of super type $A$ and of type $\mathfrak{br}(2;5)$ in Section \ref{section:examples}.
\subsection{\hspace{0.1pt} }
Here are some questions on, and potential applications of, distinguished pre-Nichols algebras that support our interest
on them.
\subsubsection{} If the diagonal braiding $(V,c)$ is given by a symmetric matrix, then the algebra $Z(V)$ is commutative.
Let $G(V)$ be the algebraic group $\Spec Z(V)$. There exists a correspondence
between (almost all) braidings of diagonal type whose Nichols algebras are finite-dimensional,
and finite-dimensional contragredient Lie superalgebras in positive characteristic \cite{AA-WGCLSandNA}.
The group $G(V)$ should be related with the Lie superalgebra corresponding to $(V,c)$;
it should allow a geometric approach to the representation theory of $U(V)$ as in \cite{DP-quantum}.
\subsubsection{} The Hopf algebra $U(V)$ is the `quantum' analogue of the enveloping Lie superalgebra in the previous point
while $\mathfrak{u}(V)$ is the one of its restricted enveloping superalgebra.
Therefore, as for Lie algebras in positive characteristic,
we can start by studying the representation theory of the algebra $U(V)$ and pass to $\mathfrak{u}(V)$.
\subsection*{Acknowledgements}
I thank Nicol\'as Andruskiewitsch for interesting and guiding discussions, and several comments which help to improve this work.
\section{Preliminaries}\label{section:preliminares}
\subsection{Conventions}
We work over an algebraically closed field $ \mathbf{k}$ of characteristic zero. Tensor products, algebras, coalgebras and Hopf algebras are taken over $ \mathbf{k}$. For each $\theta \in\mathbb{N}$, let $\I_\theta=\{1,2,\dots,\theta\}$; when $\theta$ is clear from the context, we simply set $\I=\I_\theta$.
Given a matrix $(q_{ij})_{1\le i,j\le\theta}\in \mathbf{k}^{\theta\times\theta}$, let $\widetilde{q_{ij}}=q_{ij}q_{ji}$, $i\neq j$.
We consider the $q$-polynomial numbers in $\mathbb{Z}[\mathbf{q}]$, $ n\in \mathbb{N}$, $0 \leq i \leq n$,
\begin{align*}
(n)_\mathbf{q} &=\sum_{j=0}^{n-1}\mathbf{q}^{j}, & (n)_\mathbf{q}!&=\prod_{j=1}^{n} (j)_\mathbf{q}, &
\binom{n}{i}_\mathbf{q} & =\frac{(n)_\mathbf{q}!}{(n-i)_\mathbf{q}!(i)_\mathbf{q}!}.
\end{align*}
$(n)_q$, $(n)_q!$, $\binom{n}{i}_q$ are the evaluations of the polynomials in $\mathbf{q}=q\in \mathbf{k}$.
Given $\mathbf{a}=(a_1,\dots,a_\theta)\in\mathbb{N}_0^\theta$, we use the notation: $\mathbf{t}^\mathbf{a}= t_1^{a_1}\cdots t_\theta^{a_\theta}$.
The \emph{height} of $\mathbf{a}$ is $\hgt(\mathbf{a})=\sum_{j=1}^\theta a_j$.
If $W=\oplus_{\mathbf{a}\in\mathbb{N}_0^\theta} W_\mathbf{a}$ is an $\mathbb{N}_0^\theta$-graded vector space, then its \emph{Hilbert series} is $ \mathfrak{H}_W = \sum_{\mathbf{a}\in\mathbb{N}_0^\theta} \dim W_\mathbf{a} \, \mathbf{t}^\mathbf{a} \in\mathbb{N}_0[[t_1,\dots,t_\theta]]$.
We denote by $\{\alpha_i\}_{1\le i\le\theta}$ the canonical basis on $\mathbb{Z}^\theta$. Given a bicharacter $\chi:\mathbb{Z}^{\theta}\times\mathbb{Z}^{\theta}\to \mathbf{k}^\times$ and $w\in\Aut(\mathbb{Z}^{\theta})$, $w^*\chi$ denotes the bicharacter
\begin{align*}
w^*\chi(\beta,\gamma)&=\chi(w^{-1}(\beta),w^{-1}(\gamma)), & \beta,\gamma & \in\mathbb{Z}^{\theta}.
\end{align*}
\medbreak
Let $H$ be a Hopf algebra with bijective antipode $\mathcal{S}$.
We use the Sweedler notation for the comultiplication, $\Delta(h)=h_{(1)}\otimes h_{(2)}$, $h\in H$, and for left $H$-comodules $X$, $\lambda(x)=x_{(-1)}\otimes x_{(0)}\in X\otimes H$, $x\in X$.
We denote by $^{H}_{H}\mathcal{YD}$ the category of left Yetter-Drinfeld modules over $H$. Recall that this is a braided tensor category, where the braiding for $M,N \in^{H}_{H}\mathcal{YD}$ is $c=c_{M,N}:M\otimes N \to N\otimes M$,
\begin{align*}
c(m\otimes n) &= m_{(-1)}\cdot n \otimes m_{(0)}, & \quad m\in M, \, &n\in N.
\end{align*}
The inverse braiding is $c^{-1}=c_{M,N}^{-1}:N\otimes M \to M\otimes N$,
\begin{align*}
c^{-1}(n\otimes m)&= m_{(0)}\otimes \mathcal{S}^{-1}(m_{(-1)})\cdot n, & \quad m\in M, \, &n\in N.
\end{align*}
In particular, if $V\in^{H}_{H}\mathcal{YD}$, then $(V,c_{V\otimes V})$ is a braided vector space.
If $(V,c)$ is of diagonal type with basis $E_1,\dots,E_\theta$ and matrix $(q_{ij})$, $\Gamma$ is a group
and $g_i\in\Gamma$, $\chi_j\in\widehat\Gamma$
are such that $\chi_j(g_i)=q_{ij}$, then $(V,c)$ can be realized as a Yetter-Drinfeld module over $H= \mathbf{k}\Gamma$ by defining the action and coaction:
\begin{align}
\lambda(E_i)&=g_i\otimes E_i, & g\cdot E_i&=\chi_i(g) E_i, & 1\le i\le\theta, \, & g\in\Gamma. \label{eq:YD diagonal}
\end{align}
For example take $\Gamma=\mathbb{Z}^\theta$, $g_i=\alpha_i$ and $\chi_j\in\widehat\mathbb{Z}^{\theta}$ such that $\chi_j(\alpha_i)=q_{ij}$ \cite{AS Pointed HA}.
Let $R$ be a Hopf algebra in $^{H}_{H}\mathcal{YD}$ with product $m$, coproduct $\underline{\Delta}$ and antipode $\mathcal{S}_R$. We use the following variation of Sweedler notation: $\underline{\Delta}(r)=r^{(1)}\otimes r^{(2)}$, $r\in R$. $R\# H$ denotes the \emph{bosonization} of $R$ by $H$; we refer to \cite[Section 1.5]{AS Pointed HA} for the definition.
The \emph{braided commutator} and the \emph{braided adjoint action} of $x$ on $y$, $x,y\in R$, are defined as follows
\begin{align*}
[x,y]_c&=m (\id_{R\otimes R}-c)(x\otimes y), \\ \ad_c x(y)&=m(m\otimes \mathcal{S}_R)(\id\otimes c)(\underline{\Delta}\otimes\id)(x\otimes y).
\end{align*}
The set of primitive elements is $\mathcal{P}(R)=\{r\in R: \underline{\Delta}(r)=r\otimes 1+1\otimes r \}$.
Notice that $\ad_c x(y)=[x,y]_c$ if $x\in\mathcal{P}(R)$.
\subsection{Normal coideal subalgebras and Hopf ideals}
A \emph{right coideal subalgebra} $B$ of a Hopf algebra $H$ is a subalgebra $B$ such that $\Delta(B)\subset B\otimes H$. A right coideal subalgebra $B$ is \emph{normal} if it is stable under the right adjoint action $\ad_r(h) (b)= \mathcal{S}(h_{(1)})b h_{(2)}$, $b\in B$, $h\in H$.
If $B\subset H$ is a right coideal subalgebra, then the \emph{normal right coideal subalgebra} $N(B)$\label{def:NX} is the subalgebra generated by $\{\mathcal{S}(h_{(1)})b h_{(2)}:\,h\in H,
b\in B\}$.
There is a correspondence between quotient Hopf algebras and normal right coideal subalgebras under some mild conditions.
\begin{theorem}\label{thm:takeuchi-correspondance}\cite[Theorem 3.2]{T}
The maps
\begin{align*}
&B \text{ a normal right coideal subalgebra}\rightsquigarrow {\mathcal I}(B)=H B^+, \\
&I \text{ a Hopf ideal } \rightsquigarrow {\mathcal X}(I)=H^{\co H/I},
\end{align*}
restrict to mutually inverse bijective correspondences between the set of normal
right coideal subalgebras $B$ such that $H$ is right $B$-faithfully flat and the set of Hopf ideals $I$ such that $H$ is $H/I$-coflat.\qed
\end{theorem}
There is an analogous bijection for connected Hopf algebras in ${}^H_H\mathcal{YD}$.
\begin{proposition}\cite[Proposition 3.6]{AAGMV}\label{pro:masuoka}
Let $R$ be a connected Hopf algebra in
${}^H_H\mathcal{YD}$.
\smallbreak
\noindent${\mathsf {(i)}\;}$
$R$ is a free left and right module over every right coideal subalgebra, and is a cofree left and right comodule over every quotient left module coalgebra.
\smallbreak
\noindent${\mathsf {(ii)}\;}$
The maps $B \mapsto R/RB^+$, $T \mapsto {}^{\co T}R$ give a bijection between the
set of right coideal subalgebras $B$ of $R$ and the set of quotient left
$R$-module coalgebras $T$ of $R$.
\smallbreak
\noindent${\mathsf {(iii)}\;}$
If $B$ and $T$ correspond to each other via the bijection in ${\mathsf {(ii)}\;}$, then there
exists a left $T$-colinear and right $B$-linear isomorphism $T\otimes B \simeq R$.
\qed
\end{proposition}
\subsection{Weyl groupoids and generalized root systems}\label{subsection:weylgrupoide}
We recall the notion of generalized root systems using the notation in \cite{AA-WGCLSandNA}, see also \cite{CH1}.
\subsubsection{Basic data}
Given $\I=\I_\theta$, $\mathcal{X}\neq\emptyset$ and $\rho: \I \to \Sb_{\mathcal{X}}$, the pair $(\mathcal{X}, \rho)$ is called a \emph{basic datum}
of rank $\vert\mathcal{X}\vert$ and type $\theta$ if $\rho_i^2 = \id$ for all $i\in \I$.
Let $\mathcal{Q}_{\rho}$ be the quiver $\{\sigma_i^x := (x, i, \rho_i(x)): i\in \I, x\in \mathcal{X}\}$ over $\mathcal{X}$, with target and source
$t(\sigma_i^x) = x$, $s(\sigma_i^x) = \rho_i(x)$.
In any quotient of the free groupoid $F(\mathcal{Q}_{\rho})$, we adopt the convention
\begin{align}\label{eq:conventio}
\sigma_{i_1}^x\sigma_{i_2}\cdots \sigma_{i_t} =
\sigma_{i_1}^x\sigma_{i_2}^{\rho_{i_1}(x)}\cdots \sigma_{i_t}^{\rho_{i_{t-1}} \cdots \rho_{i_1}(x)}.
\end{align}
That is, the implicit superscripts are the only possible to have compositions.
\subsubsection{Coxeter groupoids}
A \emph{Coxeter datum} is a triple $(\mathcal{X}, \rho, \mathbf{M} )$, where $(\mathcal{X}, \rho)$ is a basic datum of type $\I$
and $\mathbf{M} = (\mathbf{m} ^x)_{x\in \mathcal{X}}$, $\mathbf{m} ^x = (m^x_{ij})_{i,j\in \I}$, is a family of Coxeter matrices such that
\begin{align}\label{eq:coxeter-datum}
s((\sigma^x_i\sigma_j)^{m^x_{ij}}) &= x,& i,j&\in \I,& x&\in \mathcal{X}.
\end{align}
The \emph{Coxeter groupoid} $\mathcal{W}(\mathcal{X}, \rho, \mathbf{M} )$ \cite[Definition 1]{HY} is the groupoid generated by $\mathcal{Q}_{\rho}$ with relations
\begin{align}\label{eq:def-coxeter-gpd}
(\sigma_i^x\sigma_j)^{m^x_{ij}} &= \id_x, &i, j\in \I,\, &x\in \mathcal{X}.
\end{align}
In particular, if $i = j$, then \eqref{eq:def-coxeter-gpd} means that either $\sigma_i^x$ is an involution when $ \rho_i(x) = x$, or else that $\sigma_i^x$ is the inverse arrow of
$\sigma_i^{\rho_i}(x)$ when $\rho_i(x) \neq x$.
\subsubsection{Generalized root systems}
Let $\mathcal{C}=(C^x)_{x\in \mathcal{X}}$ be a family of generalized Cartan matrices $C^x=(c_{ij}^x)_{i,j\in\I}$ with row invariance
\begin{align}\label{eq:condicion Cartan scheme}
c^x_{ij}&=c^{\rho_i(x)}_{ij} & \mbox{for all }&x \in \mathcal{X}, \, i,j \in \I.
\end{align}
We define $s_i^x\in GL_{\theta}(\mathbb{Z})$ by
\begin{align}\label{eq:reflection-x}
s_i^x(\alpha_j)&=\alpha_j-c_{ij}^x\alpha_i, & j&\in \I,& i &\in \I, x \in \mathcal{X}.
\end{align}
Then \eqref{eq:condicion Cartan scheme} means that $s_i^x$ is the inverse of $s_i^{\rho_i(x)}$. A \emph{generalized root system} (GRS for short)
\cite[Definition 1]{HY} is a collection $\mathcal{R}:= \mathcal{R}(\mathcal{X}, \rho, \mathcal{C}, \Delta)$, where $(\mathcal{X}, \rho)$
is a basic datum of type $\I$, $\mathcal{C}$ is as above,
and $\Delta = (\Delta^x)_{x\in \mathcal{X}}$ is a family of subsets $\Delta^x \subset \mathbb{Z}^{\I}$ such that for all $x \in \mathcal{X}$, $i \neq j \in \I$,
\begin{align}
\label{eq:def root system 1}
\Delta^x &= \Delta^x_+ \cup \Delta^x_-, & \Delta^x_+ &:= \Delta^x \cap \mathbb{N}_0^I, \,\, \Delta^x_-:=-\Delta^x_+;
\\ \label{eq:def root system 2}
\Delta^x \cap \mathbb{Z} \alpha_i &= \{\pm \alpha_i \};& &
\\ \label{eq:def root system 3}
s_i^x(\Delta^x)&=\Delta^{\rho_i(x)};&&
\\ \label{eq:def root system 4}
(\rho_i\rho_j)^{m_{ij}^x}(x)&=(x), & m_{ij}^x&:=|\Delta^x \cap (\mathbb{N}_0\alpha_i+\mathbb{N}_0 \alpha_j)|.
\end{align}
We call $\Delta^x_+$, $ \Delta^x_-$ the set of \emph{positive}, respectively \emph{negative}, roots.
Let $\mathbf{M} =(M^x)_{x\in\mathcal{X}}$, $M^x=(m_{ij}^x)_{i,j\in\I}$.
The \emph{Weyl groupoid} of $\mathcal{R}$ is $\mathcal{W} = \mathcal{W}(\mathcal{X}, \rho, \mathbf{M} )$. We can describe this groupoid by using \cite[Theorem 1]{HY}.
Let $\mathcal{G} = \mathcal{X} \times GL_{\theta}(\mathbb{Z}) \times \mathcal{X}$, $\varsigma_i^x = (x, s_i^x,\rho_i(x))$, $i \in \I$, $x \in \mathcal{X}$,
and $\mathcal{W}' = \mathcal{W}(\mathcal{X}, \rho, \mathcal{C})$ the subgroupoid of $\mathcal{G}$ generated by all the $\varsigma_i^x$. There exists a morphism of quivers
$\mathcal{Q}_{\rho} \to \mathcal{G}$, $\sigma_i^x \mapsto \varsigma_i^x$ with image $\mathcal{W}'$. This induces an
isomorphism of groupoids $\mathcal{W}\to \mathcal{W}(\mathcal{X}, \rho, \mathcal{C})$.
If $w = \sigma_{i_1}^x \cdots \sigma_{i_m}$ and $\alpha\in\mathbb{Z}^{\theta}$, then we define $w(\alpha) = s_{i_1}^x \cdots s_{i_m}(\alpha)$, so
$w(\Delta^x)= \Delta^y$, by \eqref{eq:def root system 3}. The set of \emph{real roots} at $x$ is
\begin{align}
\label{defrealroot} (\Delta^{\re})^x &= \bigcup_{y\in \mathcal{X}}\{ w(\alpha_i): \ i \in \I, \ w \in \mathcal{W}(y,x) \}.
\end{align}
The \emph{length} of $w\in \mathcal{W}(x,\mathcal{X})$ is
$$ \ell(w)= \min \{ m\in \mathbb{N}_0: \ \exists i_1, \ldots, i_n \in \I \mbox{ such that }w = \sigma_{i_1}^x \cdots \sigma_{i_m} \}. $$
An expression $w = \sigma_{i_1}^x \cdots \sigma_{i_m}$ is \emph{reduced} if $m = \ell(w)$.
\begin{lemma}{\cite[Corollary 3]{HY}} \label{Lemma:longitudHY}
Let $m \in \mathbb{N}$, $x, y \in \mathcal{X}$, $i_1, \ldots, i_m,j \in I$, $w=\sigma_{i_1}^x \cdots \sigma_{i_m} \in \Hom(Y,X)$, where $\ell (w)=m$. Then
\begin{itemize}
\item $\ell (w \sigma_j)=m+1$ if and only if $w(\alpha_j) \in \Delta^x_+$,
\item $\ell (w \sigma_j)=m-1$ if and only if $w(\alpha_j) \in \Delta^x_-$.\qed
\end{itemize}
\end{lemma}
\begin{proposition}{\cite[Prop. 2.12]{CH1}} \label{Prop:maxlongitudCH}
If $w=\sigma_{i_1}^x \cdots \sigma_{i_N}\in\mathcal{W}$ is such that
$\ell(w)=N$, then all the roots $\beta_j=s_{i_1}^x\cdots
s_{i_{j-1}}(\alpha_{i_j})\in\Delta^x$ are positive and pairwise
different. In particular, if $\mathcal{R}$ is finite and $w$ is an element of maximal
length, then $\Delta^x_+= \{\beta_j | 1 \leq j\leq N\}$.
Hence $\Delta^x=(\Delta^{\re})^x$. \qed
\end{proposition}
\subsubsection{GRS of Nichols algebras with finite sets of roots}
Let $V$ be a braided vector space of diagonal type with matrix $(q_{jk})$. Let $\chi:\mathbb{Z}^{\theta}\times\mathbb{Z}^{\theta}\to \mathbf{k}^\times$ be the bicharacter such that $\chi(\alpha_j,\alpha_k)=q_{jk}$ for all $j,k\in\I$.
The set $\Delta^{V}_+$ of degrees of a PBW basis of $\mathcal{B}(V)$, counted with their multiplicities as in \cite{H-Weyl grp}, does not depend on the PBW
basis. For each $i\in \I$ let $c_{ii}^{V}=2$,
\begin{align}
-c_{ij}^{V}&:= \min \left\{ n \in \mathbb{N}_0: (n+1)_{q_{ii}}
(1-q_{ii}^n q_{ij}q_{ji} )=0 \right\} & & \notag \\
&=\max\{ n\in\mathbb{N}_0: n\alpha_i+\alpha_j\in \Delta_+^{V}\}, & j & \neq i \label{defn:mij}
\end{align}
$s_i^{V}\in\Aut(\mathbb{Z}^\theta)$ such that $s_i^{V}(\alpha_j)=\alpha_j-c_{ij}^{V}\alpha_i$. Let $\mathcal{X}$ be the set of braided vector spaces $(V,c)$ of diagonal type such that $\Delta_+^{V}$ is finite.
Set $\Delta^{V}=\Delta^{V}_+\cup (-\Delta^{V}_+)$. Define $\rho_i(V)$ as the braided vector space with associated bicharacter $\rho_i(\chi)(\alpha,\beta)= \chi(s_i^{V}(\alpha),s_i^{V}(\beta))$, $\alpha, \beta \in\mathbb{Z}^{\theta}$. $\mathcal{R}:= \mathcal{R}(\mathcal{X}, \rho, \mathcal{C}, (\Delta^x)_{x\in\mathcal{X}})$ is the GRS attached to $\mathcal{X}$. Indeed
$\Delta^{\rho_i(V)}_+ = s_i\left( \Delta^{V}_+ \setminus\{\alpha_i\}\right)
\cup \{\alpha_i\}$ by
\cite{H-Weyl grp}, so \eqref{eq:def root system 3} holds, and \cite[Theorems 6.2, 6.9]{HS-coideal subalg} completes the proof.
\subsection[Lusztig Isomorphisms]{Lusztig Isomorphisms of Nichols algebras of diagonal type}\label{section:Lusztig isomorphisms}
We recall now Lusztig type isomorphisms \cite{H-isom} of Hopf algebras related with Nichols algebras of diagonal type. They can be thought of as generalizations of the isomorphisms of quantized enveloping algebras in \cite{L-libro}.
\smallbreak
Fix a bicharacter $\chi : \mathbb{Z}^{\theta} \times \mathbb{Z}^{\theta} \to \mathbf{k}^{\times}$, $q_{ij}=\chi(\alpha_i, \alpha_j)$ and $(V,c)$ the braided vector space of diagonal type with braiding matrix $(q_{ij})$ for a basis $(E_i)_{1\le i\le\theta}$. \emph{From now on we assume that $V\in\mathcal{X}$; in particular all the $c_{ij}^{V}$ do exist}. Set $(V,c)$ as an element of $^{ \mathbf{k}\mathbb{Z}^{\theta}}_{ \mathbf{k}\mathbb{Z}^{\theta}}\mathcal{YD}$ as in \eqref{eq:YD diagonal}, so $T(V)$ is a Hopf algebra in this category. We recall definitions and results from \cite[Section 4.1]{H-isom}.
Let $\mathcal{U}(V)$ be the algebra generated by elements $E_i$,
$F_i$, $K_i$, $K_i^{-1}$, $L_i$, $L_i^{-1}$, $1 \leq i \leq \theta$,
and relations
\begin{align*}
XY&=YX, & X,Y \in & \{ K_i^{\pm}, L_i^{\pm}: 1 \leq i \leq \theta \}, \\
K_iK_i^{-1}&=L_iL_i^{-1}=1, & E_iF_j-F_jE_i&=\delta_{i,j}(K_i-L_i).
\\ K_iE_j&=q_{ij}E_jK_i, & L_iE_j&=q_{ji}^{-1}E_jL_i,
\\ K_iF_j&=q_{ij}^{-1}F_jK_i, & L_iF_j&=q_{ji}F_jL_i.
\end{align*}
If $(W,c)$ denotes the braided vector space of diagonal type corresponding to the transpose of the matrix $(q_{ij})_{i,j\in\I}$,
then $\mathcal{U}(V)$ is the quantum double of $T(V)\# \mathbf{k}\mathbb{Z}^{\theta}$ and $T(W)\# \mathbf{k}\mathbb{Z}^{\theta}$,
see \cite[Proposition 4.6]{H-isom}.
$\mathcal{U}(V)$ admits a Hopf algebra structure, where the comultiplication satisfies
\begin{align*}
\Delta(K_i^{\pm1})&=K_i^{\pm1} \otimes K_i^{\pm1}, & \Delta(E_i)&=E_i \otimes 1 + K_i \otimes E_i,
\\ \Delta(L_i^{\pm1})&=L_i^{\pm1} \otimes L_i^{\pm1}, & \Delta(F_i)&=F_i \otimes L_i + 1 \otimes F_i.
\end{align*}
Then $\mathcal{U}(V)$ is a $\mathbb{Z}^{\theta}$-graded Hopf algebra such that
\begin{align*}
\deg(K_i)&=\deg(L_i)=0, & \deg(E_i)&=\alpha_i, & \deg(F_i)&=-\alpha_i.
\end{align*}
We fix the following notation:
\begin{itemize}
\item $\mathcal{U}^+(V)$ (respectively, $\mathcal{U}^-(V)$) is the subalgebra
generated by $E_i$ (respectively, $F_i$), $1 \leq i \leq \theta$;
\item $\mathcal{U}^{\ge0}(V)$ (respectively, $\mathcal{U}^{\le0}(V)$) is the subalgebra
generated by $E_i$, $K_i$, $K_i^{-1}$ (respectively, $F_i$, $L_i$, $L_i^{-1}$), $1
\leq i \leq \theta$,
\item $\mathcal{U}^{+0}(V)$ (respectively, $\mathcal{U}^{-0}(V)$) is the subalgebra
generated by $K_i$, $K_i^{-1}$ (respectively, $L_i$, $L_i^{-1}$), $1
\leq i \leq \theta$, which are isomorphic to $ \mathbf{k} \mathbb{Z}^{\theta}$ as Hopf algebras;
\item $\mathcal{U}^0(V)$ is the subalgebra
generated by $K_i$, $K_i^{-1}$, $L_i$ and $L_i^{-1}$, which is isomorphic to $ \mathbf{k} \mathbb{Z}^{2\theta}$ as Hopf algebras.
\end{itemize}
Then $\mathcal{U}^+(V)$ is isomorphic to $T(V)$ and $\mathcal{U}^{\ge 0}(V)$ is isomorphic to $T(V)\# \mathbf{k}\mathbb{Z}^{\theta}$ as Hopf algebras.
For each homogeneous element $E\in\mathcal{U}^+(V)_n$ and $k\in\{0,1,\ldots,n\}$,
$\underline\Delta_{n-k,k}(E)$ is the component of $\underline \Delta(E)$ in $\mathcal{U}^+(V)_{n-k}\otimes\mathcal{U}^+(V)_k$.
We consider some skew-derivations as in \cite[4.2]{H-isom}.
\begin{proposition}\label{prop:derivadas torcidas}
There exist linear endomorphisms $\partial^K_i$, $\partial^L_i$ of $\mathcal{U}^+(V)$, $i\in\I$, such that for each $E\in \mathcal{U}^+(V)_n$, $n\in\mathbb{N}$,
\begin{align*}
\underline \Delta_{n-1,1}(E)&=\sum_{i=1}^\theta \partial_i^K(E) \otimes E_i, & \underline \Delta_{1,n-1}(E)&=\sum_{i=1}^\theta E_i \otimes\partial_i^L(E),
\end{align*}
Moreover, $EF_i-F_iE = \partial^K_i(E)K_i- L_i \partial^L_i(E)$ and
\begin{align*}
\partial_i^K(1)&= \partial_i^L(1)=0, & \partial_i^K(EE')&= \partial_i^K(E)(K_i \cdot E')+ E \partial_i^K(E'),
\\ \partial_i^K(E_j)&= \partial_i^L(E_j)=\delta_{i,j}, & \partial_i^L(EE')&= \partial_i^L(E)E'+ (L_i^{-1}\cdot E) \partial_i^L(E').
\end{align*}
for all $E,E' \in \mathcal{U}^+(V)$, $j\in\I$. \qed
\end{proposition}
\begin{corollary}\label{coro:FE=EF+cosas}
Let $m,n\in\mathbb{N}$, $E\in\mathcal{U}^+(V)_m$, $F\in\mathcal{U}^-(V)_n$.
\smallbreak
\noindent${\mathsf {(i)}\;}$ If $m\geq n$, then $FE\in EF+\sum_{j=0}^{n-1} \mathcal{U}^+(V)_{m-n+j}\mathcal{U}^0(V)\mathcal{U}^-(V)_j$.
\smallbreak
\noindent${\mathsf {(ii)}\;}$ If $m<n$, then $FE\in EF+\sum_{j=0}^{m-1} \mathcal{U}^+(V)_{j}\mathcal{U}^0(V)\mathcal{U}^-(V)_{n-m+j}$.
\end{corollary}
\begin{proof}
${\mathsf {(i)}\;}$ We argue by induction on $n$. If $n=1$, then $F$ is a linear combination of $F_i$'s and Proposition \ref{prop:derivadas torcidas} applies.
Let $n>1$. We may assume that $F=F_iF'$, $F'\in\mathcal{U}^+(V)_{n-1}$. By inductive hypothesis,
$$ FE=F_iF'E\in F_iEF'+\sum_{j=0}^{n-2} F_i\mathcal{U}^+(V)_{m-n+1+j}\mathcal{U}^0(V)\mathcal{U}^-(V)_j,$$
and the proof follows by Proposition \ref{prop:derivadas torcidas}.
\smallbreak
\noindent${\mathsf {(ii)}\;}$ We may assume that $F=F_{i_1}\dots F_{i_{n-m}} F'$, $F'\in\mathcal{U}^+(V)_m$. By ${\mathsf {(i)}\;}$,
\begin{align*}
FE=F_{i_1}\dots F_{i_{n-m}}F'E\in & F_{i_1}\dots F_{i_{n-m}}EF'\\
&+\sum_{j=0}^{m-1} F_{i_1}\dots F_{i_{n-m}} \mathcal{U}^+(V)_{j}\mathcal{U}^0(V)\mathcal{U}^-(V)_j,
\end{align*}
and the proof follows again by Proposition \ref{prop:derivadas torcidas}.
\end{proof}
\bigbreak
Fix $i\in\I$. For each $j\neq i$ we set as in \cite{H-isom}:
\begin{align*}
E_{j,0(i)}^+=E_{j,0(i)}^-&=E_j, & F_{j,0(i)}^+=F_{j,0(i)}^-&=F_j,
\end{align*}
and recursively,
\begin{gather}\label{eq:def Ejm+,Fjm+}
\begin{aligned}
E_{j,m+1(i)}^+ &:= E_iE_{j,m(i)}^+ - (K_i \cdot E_{j,m(i)}^+)E_i = (\ad_c E_i)^{m+1}E_j, \\
F_{j,m+1(i)}^+ &:= F_iF_{j,m(i)}^+ - (L_i \cdot F_{j,m(i)}^+)F_i= (\ad_c F_i)^{m+1}F_j,
\\ E_{j,m+1(i)}^- &:= E_iE_{j,m(i)}^- - (L_i \cdot E_{j,m(i)}^-)E_i,
\\ F_{j,m+1(i)}^- &:= F_iF_{j,m(i)}^- - (K_i \cdot F_{j,m(i)}^-)F_i.
\end{aligned}
\end{gather}
When $i$ is explicit, we simply denote $E_{j,m(i)}^\pm$ by $E_{j,m}^\pm$.
\begin{remark} \cite[Lemma 4.23, Corollary 4.25]{H-isom}
For all $N\in\mathbb{N}$,
\begin{align}
E_{j,N}^+ &= \sum_{s=0}^N (-1)^s q_{ij}^s q_{ii}^{s(s-1)/2} \binom{N}{s}_{q_{ii}} E_i^{N-s}E_jE_i^s,
\label{eq:expresion adjunta} \\
E_i^N F_i-F_iE_i^N&= (N)_{q_{ii}}(q_{ii}^{1-N}K_i-L_i)E_i^{N-1}, \label{eq:simple power root contra Fp} \\
E_{j,N}^+ F_i - F_i E_{j,N}^+&= (N)_{q_{ii}}(q_{ii}^{N-1}\widetilde{q_{ij}}-1)L_i E_{j,N-1}^+. \label{eqn:corchete E+ con Fp}
\end{align}
The coproduct satisfies the following identities:
\begin{align}
\underline\Delta \left( E_i^N \right) &= \sum_{s=0}^N \binom{N}{s}_{q_{ii}} E_i^{s} \otimes E_{i}^{N-s},
\label{eq:coproducto Ep a la N} \\
\underline\Delta \left( E_{j,N}^+ \right) &= E_{j,N}^+ \otimes 1 + \sum_{s=0}^{N-1} \left( \prod_{r=N-s}^{N-1} (r)_{q_{ii}}(1-q_{ii}^{r}\widetilde{q_{ij}}) \right) E_i^{s} \otimes E_{j,N-s}^+,
\label{eq:coproducto Ei+} \\
\underline\Delta \left( E_{j,N}^- \right) &= 1 \otimes E_{j,N}^- \notag \\
& \quad + \sum_{s=0}^{N-1} q_{ij}^s\left( \prod_{r=N-s}^{N-1} (r)_{q_{ii}}\left(1-q_{ii}^{-r}\widetilde{q_{ij}}^{-1}\right) \right) E_{j,N-s}^- \otimes E_i^{s}.
\label{eq:coproducto Ei-}
\end{align}
\end{remark}
Let $\mathcal{J}^{\pm}(V)$ be the ideal of $\mathcal{U}^\pm(V)$ such that $\mathfrak{u}^+(V)=\mathcal{U}^+(V)/\mathcal{J}^+(V)$, $\mathfrak{u}^-(V)=\mathcal{U}^-(V)/\mathcal{J}^-(V)$ are isomorphic to $\mathcal{B}(V)$, $\mathcal{B}(W)$, respectively, and set $\mathfrak{u}(V)= \mathcal{U}(V) / (\mathcal{J}^-(V)+\mathcal{J}^+(V))$.
Then $\mathfrak{u}(V)$ is the quantum double of the algebras $\mathcal{B}(V)\# \mathbf{k}\mathbb{Z}^\theta$ and $\mathcal{B}(W)\# \mathbf{k}\mathbb{Z}^\theta$ by \cite[Theorem 5.8]{H-isom}.
We need another quotients of $\mathcal{U}(V)$ in order to introduce Lusztig isomorphisms. First we recall
\cite[Definition 2.6]{A-presentation}.
An element $i\in\I$ is a \emph{Cartan vertex} of $V$ if $\widetilde{q_{ij}}=q_{ii}^{c_{ij}^{V}}$ for all $j \neq i$.
The set of \emph{Cartan roots} is
\begin{equation}\label{eq:def O(chi)}
\mathcal{O}(V):=\{ s_{i_1}^{V}\dots s_{i_k}(\alpha_i): i \mbox{ is a Cartan vertex of } \rho_{i_k}\dots\rho_{i_1}(V)\}.
\end{equation}
Set $N_i= \ord q_{ii}$.
$\mathcal{J}_i^+(V)$, $\mathcal{J}_i^-(V)$ are the ideals of $\mathcal{U}^+(V)$, respectively $\mathcal{U}^-(V)$, generated by
\begin{itemize}
\item[(a)] $E_i^{N_i}$, respectively $F_i^{N_i}$, if $i$ is not a Cartan vertex,
\item[(b)] $E_{j,-c_{ij}^{V}+1}^+$, respectively $F_{j,-c_{ij}^{V}+1}^+$, for each $i$ such that $N_i\geq-c_{ij}^{V}+1$.
\end{itemize}
Set also
\begin{align*}
\mathcal{U}_i(V)&:=\mathcal{U}(V)/\left(\mathcal{J}_i^+(V)+\mathcal{J}_i^-(V) \right), & \mathcal{U}_i^\pm(V)&:=\mathcal{U}^\pm(V)/\mathcal{J}_i^\pm(V).
\end{align*}
Let $\underline E_j$, $\underline F_j$, $\underline K_j$, $\underline L_j$ be the generators of $\mathcal{U}({\rho_i(V)})$. Set
\begin{equation}\label{eqn:escalares lambda}
\lambda_j(V):= (-c_{ij}^{V})_{q_{ii}}\prod_{s=0}^{-c_{ij}^{V}-1}(q_{ii}^s\widetilde{q_{ij}}-1) , \qquad j \neq i.
\end{equation}
\begin{theorem}{\cite[Lemma 6.5, Theorem 6.12]{H-isom}}\label{thm: iso Lusztig-Heck}
There exist algebra maps
\begin{equation}\label{eqn:iso lusztig para cUp}
T_i, T_i^-: \mathcal{U}(V) \to \mathcal{U}_i({\rho_i(V)})
\end{equation}
univocally determined by the following conditions:
\begin{align*}
T_i(K_i)&=T_i^-(K_i)=\underline K_i^{-1}, & T_i(K_j)&=T_i^-(K_j)=\underline K_i^{-c_{ij}^{V}}\underline K_j,
\\ T_i(L_i)&=T_i^-(L_i)=\underline L_i^{-1}, & T_i(L_j)&=T_i^-(L_j)=\underline L_i^{-c_{ij}^{V}}\underline L_j,
\\ T_i(E_i)&=\underline F_i\underline L_i^{-1}, & T_i(E_j)&=\underline E^+_{j,-c_{ij}^{V}},
\\ T_i(F_i)&=\underline K_i^{-1}\underline E_i, & T_i(F_j)&=\lambda_i({\rho_i(V)})^{-1}\underline F^+_{j,-c_{ij}^{V}},
\\ T_i^-(E_i)&=\underline K_i^{-1}\underline F_i, & T_i^-(E_j)&=\lambda_i({\rho_i(V)}^{-1})^{-1}\underline E^-_{j,-c_{ij}^{V}},
\\ T_i^-(F_i)&=\underline E_i\underline L_i^{-1}, & T_i^-(F_j)&=\underline F^-_{j,-c_{ij}^{V}}.
\end{align*}
Such morphisms induce algebra isomorphisms (denoted by the same name):
\begin{align}
T_i, T_i^-: \, &\mathfrak{u}(V) \to \mathfrak{u}(\rho_i(V)) & \mbox{such that }& T_iT_i^-=T_i^-T_i=\id.\qed
\end{align}
\end{theorem}
\subsection{Lusztig isomorphisms and PBW bases}\label{subsection:pbw}
Fix an reduced expression $w=\sigma_{i_1}^{V} \sigma_{i_2}\cdots \sigma_{i_M}$ of the element of maximal length of
$\mathcal{W}(V)$. If $1\leq k\leq M$ then set $\beta_k:= s_{i_1}^{V}\cdots s_{i_{k-1}}(\alpha_{i_k})$,
$q_k:=\chi(\beta_k,\beta_k)$, $N_k = \ord q_k\in\mathbb{N}\cup\{\infty\}$.
By Proposition \ref{Prop:maxlongitudCH} $\beta_k\neq\beta_l$ if $k\neq l$, and $\Delta_+^{V}=\{\beta_k|1\leq k\leq M\}$.
Let
\begin{align}\label{eq:PBW generators}
E_{\beta_k}&=T_{i_1}\cdots T_{i_{k-1}}(E_{i_k})\in\mathfrak{u}(V)^+_{\beta_k}, & F_{\beta_k}&=T_{i_1}\cdots T_{i_{k-1}}(F_{i_k})\in\mathfrak{u}(V)^-_{\beta_k}.
\end{align}
and for each $\mathbf{a}=(a_1,\dots,a_M)\in\mathbb{N}_0^M$,
\begin{align}\label{eq:def E a la a, F a la a}
\mathbf E^{\mathbf{a}}&=E_{\beta_M}^{a_M}E_{\beta_{M-1}}^{a_{M-1}} \cdots
E_{\beta_1}^{a_1}, & \mathbf F^{\mathbf{a}}&= F_{\beta_M}^{a_M}F_{\beta_{M-1}}^{a_{M-1}} \cdots
F_{\beta_1}^{a_1}.
\end{align}
\begin{theorem} \label{thm: HY PBW bases}
\cite[Theorems 4.5, 4.8, 4.9]{HY-shapov}
The sets
\begin{align*}
& \{ \mathbf E^{\mathbf{a}}\, | \, \mathbf{a}\in\mathbb{N}_0^M, \, 0\leq a_k < N_k, \, 1\leq k\leq
M\},
\\ & \{ \mathbf F^{\mathbf{a}} \, | \, \mathbf{a}\in\mathbb{N}_0^M, \, 0\leq a_k < N_k, \, 1\leq k\leq
M\},
\end{align*}
are bases of the vector spaces $\mathfrak{u}^+(V)$, $\mathfrak{u}^-(V)$, respectively.\qed
\end{theorem}
\section[Distinguished pre-Nichols algebras]{Distinguished pre-Nichols algebras}
\label{section:Distinguished pre-Nichols algebras}
From now on we assume that $\mathcal{B}(V)$ \emph{is finite-dimensional}.
\subsection{\hspace{1pt}} We now consider some intermediate quotients between $\mathcal{U}(V)$ and $\mathfrak{u}(V)$.
\begin{definition}
Let $\mathcal{I}(V)$ be the ideal of $T(V)$ generated by all
the relations in \cite[Theorem 3.1]{A-presentation}, except the power root vectors $E_\alpha^{N_\alpha}$, $\alpha\in\mathcal{O}(V)$,
plus the quantum Serre relations $(\ad_c E_i)^{1-c_{ij}^{V}} E_j$ for those $i\neq j$ such that
$q_{ii}^{c_{ij}^{V}}=q_{ij}q_{ji}=q_{ii}$.
Then $\widetilde{\mathcal{B}}(V)=T(V)/\mathcal{I}(V)$ is the \emph{distinguished pre-Nichols algebra} of $(V,c)$.
\end{definition}
We identify $\mathcal{I}(V)$ as an ideal $\mathcal{I}^+(V)$ of $\mathcal{U}^+(V)$. Let $\mathcal{I}^-(V)$ be the corresponding ideal of $\mathcal{U}^-(V)$.
We denote by $U(V)$ the quotient of $\mathcal{U}(V)$ by the ideal generated by $\mathcal{I}^+(V)$ and $\mathcal{I}^-(V)$. By abuse of notation we denote by $E_i$, $F_i$, $K_i^\pm$, $L_i^{\pm}$ the generators of $U(V)$.
$U^+(V)=\widetilde{\mathcal{B}}(V)$, $U^-(V)$ are, respectively, the subalgebras of $U(V)$ generated by $E_i$, $F_i$. Let $U^0(V)$ be the subalgebra generated by $K_i^\pm$, $L_i^{\pm}$.
\begin{proposition}\cite[Proposition 3.3]{A-presentation}\label{prop:proyeccion sobre Nichols}
$U(V)$ is a Hopf algebra and there exist
a canonical Hopf algebra morphism $\pi_V:U(V)\twoheadrightarrow\mathfrak{u}(V)$ such that
$\pi_V\left(U^\pm(V)\right)=\mathfrak{u}^\pm(V)$. The multiplication $ m:U^+(V)\otimes U^0(V)\otimes U^-(V)\to U(V) $
gives an isomorphism of graded vector spaces. \qed
\end{proposition}
$U(V)$ is the quantum double of $\widetilde{\mathcal{B}}(V)\# \mathbf{k}\mathbb{Z}^{\theta}$ and $\widetilde{\mathcal{B}}(W)\# \mathbf{k}\mathbb{Z}^{\theta}$ since there exists a Hopf pairing induced by the one between $T(V)\# \mathbf{k}\mathbb{Z}^{\theta}$ and $T(W)\# \mathbf{k}\mathbb{Z}^{\theta}$.
\begin{remark}
Let $\mathfrak g$ be a finite-dimensional semisimple Lie algebra over $\mathbb{C}$, $C=(c_{ij})\in\mathbb{Z}^{\I\times\I}$ its finite Cartan matrix and $D=\operatorname{diag}(d_1,\dots,d_\theta)$ such that $DA$ is symmetric. If $q$ is a root of unity of odd order, then the symmetric matrix $(q_{ij})_{i,j\in\I}$, $q_{ij}=q^{d_ic_{ij}}$, defines a braiding of Cartan type. The small quantum group $\mathfrak{u}_q(\mathfrak g)$ is the quotient of $\mathfrak{u}(V)$ by the central elements $K_i-L_i^{-1}$, $1\le i\le\theta$, while the quantized enveloping algebra $U_q(\mathfrak g)$ is obtained from $U(V)$ analogously.
\end{remark}
The Lusztig isomorphisms descend to the family of algebras $U(V)$.
\begin{proposition}\cite[Proposition 3.26]{A-presentation}\label{prop:iso Lusztig para las U}
The maps \eqref{eqn:iso lusztig para cUp} induce algebra isomorphisms $ T_i, T_i^-: U(V) \to U({\rho_i(V)})$
such that $T_i T_i^-=T_i^- T_i= \id_{U(V)}$.\qed
\end{proposition}
\begin{remark}\label{rem:PRV=0 si alpha no cartan}
Using the Lusztig isomorphisms we deduce that
\begin{align*}
E_\alpha^{N_\alpha}, F_\alpha^{N_\alpha}&\neq0 \mbox{ for all }\alpha\in\mathcal{O}(V), & E_\alpha^{N_\alpha}&=F_\alpha^{N_\alpha}=0 \mbox{ for all }\alpha\notin\mathcal{O}(V),
\end{align*}
on $U(V)$. The Hilbert series of $U^\pm(V)$ is:
$$ \mathcal{H}_{U^\pm(V)}= \left( \prod_{\alpha\in \Delta_+^{V}\setminus\mathcal{O}(V) } (t^{\alpha})_{N_\alpha}\right)\bigcup \left( \prod_{\alpha\in \mathcal{O}(V) } \frac{1}{1-t^{\alpha}} \right). $$
Indeed $U^+(V)$ has a PBW basis of Lyndon hyperwords as in \cite{Kh} with the same PBW generators of $\mathfrak{u}^+(V)$, see the proof of \cite[Theorem 3.1]{A-presentation}.
\end{remark}
Set $E_\alpha, \mathbf E^{\mathbf{a}}\in U^+(V)$, $F_\alpha, \mathbf F^{\mathbf{a}}\in U^-(V)$, $\alpha\in\Delta_+^{V}$, $\mathbf{a}\in\mathbb{N}_0^M$, as in \eqref{eq:PBW generators}, \eqref{eq:def E a la a, F a la a}.
\begin{theorem} \label{thm: PBW bases preNichols}
The sets
\begin{align*}
& \{ \mathbf E^{\mathbf{a}}\, | \, \mathbf{a}\in\mathbb{N}_0^M, \, 0\leq a_k < N_k \mbox{ if }\beta_k\notin\mathcal{O}(V) \}, \\
& \{ \mathbf F^{\mathbf{a}}\, | \, \mathbf{a}\in\mathbb{N}_0^M, \, 0\leq a_k < N_k \mbox{ if }\beta_k\notin\mathcal{O}(V) \},
\end{align*}
determine bases of the vector space $U^+(V)$, $U^-(V)$, respectively.
\end{theorem}
\begin{proof}
The expression of the Hilbert series reduces the problem to the linearly independence of this set, which is proved following the same recursion as \cite[Theorem 4.5]{HY-shapov}.
\end{proof}
\subsection{\hspace{1pt}}
Let $\mathcal{B}_{i}$ be the algebra generated by $E_i$ on $U(V)$. This is a braided graded Hopf algebra. Its graded dual $\mathcal{B}_{i}^*$ is also a graded braided Hopf algebra.
There exists a projection $\pi_{i,V}:U^+(V)\twoheadrightarrow\mathcal{B}_{i}$ of braided graded Hopf algebras annihilating all the $E_j$ for $j\neq i$; the inclusion $\iota_{i,V}:\mathcal{B}_{i}\hookrightarrow U^+(V)$ is a section for this projection.
\begin{remark}\label{rem:dual algebra vertice cartan}
Let $i$ be a Cartan vertex. $\mathcal{B}_{i}^*$ has a basis $\{\mathfrak{E}_i^{(n)}:n\in\mathbb{N}_0\}$, where $\mathfrak{E}_i^{(n)}(E_i^m)=\delta_{n,m}$. The algebra and the coalgebra structures satisfy:
\begin{align}\label{eq:producto coproducto dual}
\mathfrak{E}_i^{(j)}\cdot \mathfrak{E}_i^{(k)}&= \binom{j+k}{j}_{q_{ii}}\mathfrak{E}_i^{(k+j
)}, & \underline{\Delta}(\mathfrak{E}_i^{(n)})&=\sum_{j=0}^n \mathfrak{E}_i^{(j)}\otimes \mathfrak{E}_i^{(n-j)},
\end{align}
with unit $\mathfrak{E}_i^{(0)}=1$, and counit $\varepsilon(\mathfrak{E}_i^{(n)})=\delta_{n,0}$. In particular $\mathfrak{E}_i^{(j)}\cdot \mathfrak{E}_i^{(N_i-j)}=0$ if $1\leq j\leq N_i-1$. As an algebra $\mathcal{B}_{i}^*$ is generated by $\mathfrak{E}_i^{(1)}$, $\mathfrak{E}_i^{(N_i)}$.
\end{remark}
\begin{remark}\label{rem:accion dual Bi}
There exist left and right actions of $\mathcal{B}_{i}^*$ on $U^+(V)$ given by
\begin{align}\label{eq:acciones dual}
\mathfrak{E}\triangleright Y & = Y_{(1)} \mathfrak{E} (\pi_{i,V}(Y_{(2)})), & Y \triangleleft\mathfrak{E} & = \mathfrak{E}(\pi_{i,V}(Y_{(1)}))Y_{(2)},
\end{align}
for each $\mathfrak{E}\in \mathcal{B}_{i}^*$, $Y\in U^+(V)$.
In particular,
\begin{align}\label{eq:accion X1}
\mathfrak{E}_i^{(1)}\triangleright Y & = \partial_i^K(Y), & Y \triangleleft \mathfrak{E}_i^{(1)} & =\partial_i^L(Y), & \mbox{for all }& Y\in U^+(V).
\end{align}
\end{remark}
\smallbreak
\begin{lemma}
For all $X\in U^+(V)_\beta$, $Y\in U^+(V)$, $t\in\mathbb{N}$,
\begin{equation}\label{eq:Efi como skew derivation}
(XY)\triangleleft\mathfrak{E}_i^{(t)}= \sum_{r=0}^t \chi(\beta-r\alpha_i,\alpha_i)^{t-r} \left(X\triangleleft\mathfrak{E}_i^{(r)}\right) \left(Y\triangleleft\mathfrak{E}_i^{(t-r)}\right).
\end{equation}
\end{lemma}
\begin{proof}
Notice that
\begin{align*}
\sum_{t\geq 0} E_i^t \otimes (XY)\triangleleft\mathfrak{E}_i^{(t)} & =(\pi_{i,V}\otimes\id)\underline\Delta(XY)= (\pi_{i,V}\otimes\id)\left(\underline\Delta(X)
\cdot \underline\Delta(Y)\right) \\
& = \left( \sum_{r\geq 0} E_i^r\otimes X\triangleleft\mathfrak{E}_i^{(r)} \right) \cdot \left( \sum_{s\geq 0} E_i^s\otimes Y\triangleleft\mathfrak{E}_i^{(s)} \right) \\
& = \sum_{r,s\geq 0} \chi(\beta-r\alpha_i,\alpha_i)^s \, E_i^{r+s} \otimes (X\triangleleft\mathfrak{E}_i^{(r)}) ( Y\triangleleft\mathfrak{E}_i^{(s)}).
\end{align*}
Then compare the terms of the form $E_i^t\otimes -$.
\end{proof}
\subsection{\hspace{1pt}}
Let $U^+_{\pm i}(V)$ be the subalgebra generated by $E_{j,N}^\pm$, $j\neq i$, $N\in\mathbb{N}_0$.
Given $\alpha=\sum\limits_{i=1}^\theta n_i\alpha_i\in\mathbb{Z}^{\theta}$, set $K_\alpha=\prod\limits_{i=1}^\theta K_i^{n_i}$, $L_\alpha=\prod\limits_{i=1}^\theta L_i^{n_i}\in U^0(V)$.
\begin{lemma}\label{lema:kernel derivaciones generado por ad}
${\mathsf {(i)}\;}$ $U_{+ i}^+(V)=U^+(V)^{\co \pi_{i,V}}$, $U_{-i}^+(V)=^{\co \pi_{i,V}}U^+(V)$. Then there exist isomorphisms of graded vector spaces $U^+(V) \cong U_{\pm i}^+(V) \otimes \mathcal{B}_{i}$.
\smallbreak
\noindent${\mathsf {(ii)}\;}$ If $i$ is not a Cartan vertex, then $\ker(\partial_i^K)=U_{+ i}^+(V)$, $\ker(\partial_i^L)=U_{-i}^+(V)$.
\smallbreak
\noindent${\mathsf {(iii)}\;}$ If $i$ is a Cartan vertex, then $\ker(\partial_i^K)= U_{+ i}^+(V) \mathbf{k}\left[E_i^{N_i}\right]$,
$\ker(\partial_i^L)= U_{-i}^+(V) \mathbf{k}\left[E_i^{N_i}\right]$.
\end{lemma}
\begin{proof}
The claims about $U_{+i}^+(V)$ follow by \cite[Lemma 2.4]{A-standard}, \cite[Lemma 4.31]{H-isom}, so we prove those
about $U_{-i}^+(V)$.
By \eqref{eq:coproducto Ei-} $U_{-i}^+(V)$ is a right coideal subalgebra contained in $B=^{\co \pi_{i,V}}U^+(V)$. We claim that $U_{-i}^+(V) \mathcal{B}_{i}$ is a left ideal of $U^+(V)$. Indeed if $E\in U_{-i}^+(V)$, then $E_jE\in U_{-i}^+(V)$ for $j\neq i$ since $E_j\in U_{-i}^+(V)$, and $E_i X\in U_{-i}^+(V) \mathcal{B}_{i}$ since
\begin{align*}
E_i E_{j,n}^- &= E_{j,n+1}^- + \chi(n\alpha_i+\alpha_j,-\alpha_i) E_{j,n}^- E_i, & j\neq i, \, & n\in\mathbb{N}_0.
\end{align*}
But $1\in U_{-i}^+(V) \mathcal{B}_{i}$, so $U_{-i}^+(V) \mathcal{B}_{i}=U^+(V)$. By Proposition \ref{pro:masuoka} ${\mathsf {(iii)}\;}$, there exists an isomorphism $B\otimes \mathcal{B}_{i}\simeq U^+(V)$, so $B= U_{-i}^+(V)$ and the multiplication gives this isomorphism $U_{-i}^+(V) \otimes \mathcal{B}_{i} \simeq U^+(V)$. Therefore for each $X\in U^+(V)$ there exist unique $X_n\in U_{-i}^+(V)$ such that $X= \sum_{n\geq 0}X_n E_i^n$. If $X$ has degree $\beta$, then $\partial_i^L(X) = \sum_{n\geq 1} (n)_{q_{ii}} \chi(n\alpha_i-\beta,\alpha_i) \, X_n E_i^{n-1}$ since $U_{-i}^+(V)\subseteq \ker\partial_i^L$. If $X\in\ker\partial_i^L$, then $X_n=0$ for all $n\notin\mathbb{N} N_i$.
\end{proof}
\medbreak
\begin{remark}\label{rem:algebra de Hopf YD para U+i}
$U_{+ i}^+(V)$ is a braided Hopf algebra in $^{\mathcal{B}_{i}\# \mathbf{k}\mathbb{Z}^\theta}_{\mathcal{B}_{i}\# \mathbf{k}\mathbb{Z}^\theta}\mathcal{YD}$. Indeed Lemma \ref{lema:kernel derivaciones generado por ad}${\mathsf {(iii)}\;}$ says that $U_{+ i}^+(V)=U^+(V)^{\co \pi_{i,V}}$ , and by \cite[Lemma 3.1]{AHS}
$$ U^+(V)^{\co \pi_{i,V}}=(U^+(V)\# \mathbf{k}\mathbb{Z}^\theta)^{\co \pi_{i,V}\#\id} .$$
The action and coaction satisfy, for each $E\in U^+_{+i}(V)_\beta$,
\begin{align}\label{eq:accion-coaccion a izquierda}
E_i\rightharpoonup E&=(\ad_c E_i) E, & \lambda(E)&= \sum_{n\geq 0} E_i^n K_{\beta-n\alpha_i}\otimes E\triangleleft \mathfrak{E}_i^{(n)}.
\end{align}
\end{remark}
\begin{remark}\label{rem:algebra de Hopf YD para U-i}
If $R$ is braided Hopf algebra $R$ in $^{ \mathbf{k}\mathbb{Z}^{\theta}}_{ \mathbf{k}\mathbb{Z}^{\theta}}\mathcal{YD}$, then there exists an structure of braided Hopf algebra $R^{\bop}$ as in \cite[Proposition 2.2.4]{AG} with underlying Yetter-Drinfeld module $R$ and
\begin{align*}
m^{\bop}&=m\circ c_{R,R} , & \underline{\Delta}^{\bop}&= c_{R,R}^{-1}\circ \underline{\Delta}, & \mathcal{S}^{\bop}&= \mathcal{S}.
\end{align*}
This applies for $U^+(V)$, $\mathcal{B}_{i}$, and $\pi_{i,V}:U^+(V)^{\bop}\twoheadrightarrow\mathcal{B}_{i}^{\bop}$ is a Hopf algebra map in $^{ \mathbf{k}\mathbb{Z}^{\theta}}_{ \mathbf{k}\mathbb{Z}^{\theta}}\mathcal{YD}$. Consider the Hopf algebras $U^+(V)^{\bop}\# \mathbf{k}\mathbb{Z}^\theta$, $\mathcal{B}_{i}^{\bop}\# \mathbf{k}\mathbb{Z}^\theta$, and the Hopf algebra maps $\pi_{i,V}\#\id$, $\iota_{i,V}\#\id$. Then
$$ (U^+(V)^{\bop}\# \mathbf{k}\mathbb{Z}^\theta)^{\co \pi_{i,V}\#\id}= (U^+(V)^{\bop})^{\co \pi_{i,V}}=^{\co \pi_{i,V}}U^+(V)=U_{-i}^+(V) $$
is a braided Hopf algebra in $^{\mathcal{B}_{i}^{\bop}\# \mathbf{k}\mathbb{Z}^\theta}_{\mathcal{B}_{i}^{\bop}\# \mathbf{k}\mathbb{Z}^\theta}\mathcal{YD}$.
The action and coaction satisfy
\begin{align}\label{eq:accion-coaccion para U-i}
E_i\rightharpoonup' E&=[E,E_i]_c, & \lambda'(E)&= \sum_{n\geq 0} E_i^n K_{\beta-n\alpha_i}\otimes \mathfrak{E}_i^{(n)} \triangleright E,
\end{align}
for each $E\in U^+_{-i}(V)_\beta$. But $U_{-i}^+(V)$ has the opposite product, so we take the braided Hopf algebra structure on $U_{-i}^+(V)$ obtained by applying the $\bop$ construction. The coproduct $\Delta_i:U^+_{-i}(V)\to U^+_{-i}(V)\underline{\otimes} U^+_{-i}(V)$ is given by
\begin{align}\label{eq:formula delta sub i}
\Delta_i(x) &= x_{(1)} \otimes \iota_{i,V}\pi_{i,V}( \mathcal{S}^{-1}(x_{(2)})) x_{(3)} , & x\in & U^+_{-i}(V)
\end{align}
Here $U^+_{-i}(V)\underline{\otimes} U^+_{-i}(V)$ denotes the space $U^+_{-i}(V)\underline{\otimes} U^+_{-i}(V)$ with the algebra structure viewed as an element in $^{\mathcal{B}_{i}^{\bop}\# \mathbf{k}\mathbb{Z}^\theta}_{\mathcal{B}_{i}^{\bop}\# \mathbf{k}\mathbb{Z}^\theta}\mathcal{YD}$,
\begin{align*}
(x\otimes y) \cdot (w\otimes z)
& = \sum_{n=0}^{N_i-1} \chi(\beta,\gamma-n\alpha_i)
\, x \left(\mathfrak{E}_i^{(n)} \triangleright w \right) \underline{\otimes} \left(E_i^n \rightharpoonup' y \right) z,
\end{align*}
for $x,y,w,z\in U_{-i}^+(V)$, where $y$, $w$ are homogeneous of degrees $\beta,\gamma\in\mathbb{N}_0^\theta$.
\end{remark}
\begin{remark}\label{rem:Ti(U-)=U+} By \cite[Lemma 6.7]{H-isom},
\begin{align}
q_{ii}\partial_i^L \left(T_i(X)\right) &
=-\chi(\beta,\alpha_i)^{-1} T_i(E_i \rightharpoonup' X),
\label{eq:Ti con delta L}\\
T_i \left(K_i^{-1}\cdot\partial_i^K(X)\right) &
= -\left( \underline E_\rightharpoonup T_i(X)\right),
\label{eq:Ti con delta K}\\
T_i(E_{j,n}^-)&= \underline{q}_{ii}^{\,\, n} \left(\prod_{t=-c_{ij}^{V}-n}^{-c_{ij}^{V}-1}
(t+1)_{\underline{q}_{ii}}(1-\underline{q}_{ii}^{\,\, t}\underline{q}_{ij}\underline{q}_{ji}) \right) \underline E_{j,-c_{ij}^{V}-n}^+, \label{eq:Ti(E-)=E+}
\end{align}
for each $X\in U^+_{-i}(V)_\beta$, $\beta\in\mathbb{N}_0^\theta$, $n\in\mathbb{N}_0$, where $(\underline{q}_{jk})_{j,k\in\I}$ is the matrix of $\rho_i(V)$. In particular $T_i(U^+_{-i}(V))= U^+_{+i}(\rho_i(V))$.
\end{remark}
\bigbreak
\subsection{\hspace{1pt}}
Next results resemble \cite[Lemma 4.6, Theorem 4.8]{HY-shapov}.
\begin{lemma}\label{lem:base de ker pi}
The set
\begin{align}\label{eq:base de U+i}
& \{ \mathbf E^{\mathbf{a}}\, | \, \mathbf{a}\in\mathbb{N}_0^M, \, a_1=0, \, 0\leq a_k < N_k \mbox{ if }\beta_k\notin\mathcal{O}(V) \}
\end{align}
is a basis of the vector space $U^+_{+i}(V)$.
\end{lemma}
\begin{proof}
Set $i=i_1$. By Lemma \ref{lema:kernel derivaciones generado por ad} ${\mathsf {(i)}\;}$, $E_{\beta_\ell}=\sum_{k\geq 0} X_k E_i^k$ for unique $X_k\in\ U^+_{+i}(V)$. By \eqref{eq:Ti(E-)=E+}, $T_i^-(U^+_{+i}(V))=U^+_{-i}(\rho_i(V))$, so $T_i^-(X_k)\in U^+(\rho_i(V))$. As
$$ \sum_{k\geq 0} T_i^-(X_k) (K_i^{-1}F_i)^k=T_i^-(E_{\beta_\ell})=T_{i_2}\dots T_{i_{\ell-1}}(E_{i_l})\in U^+(V),$$
last statement of Proposition \ref{prop:proyeccion sobre Nichols} implies that $X_k=0$ for all $k>0$, so $E_{\beta_\ell}=X_0\in U^+_{+i}(V)$. Then the set \eqref{eq:base de U+i} is contained in $U^+_{+i}(V)$. By Lemma \ref{lema:kernel derivaciones generado por ad} ${\mathsf {(i)}\;}$ and Theorem \ref{thm: PBW bases preNichols}, this set should also generate $U^+_{+i}(V)$ because it generates a subspace with the same Hilbert series.
\end{proof}
\begin{proposition} \label{prop: corchete entre Ebetas}
For each pair $1\leq k< \ell \leq M$,
\begin{align}
E_{\beta_k}E_{\beta_\ell}- \chi(\beta_k,\beta_\ell)E_{\beta_\ell}E_{\beta_k}
&=\sum c_{a_{k+1},\ldots,a_{\ell-1}} E_{\beta_{\ell-1}}^{a_{\ell-1}} \cdots E_{\beta_{k+1}}^{a_{k+1}} \in \mathfrak{u}^+(V), \label{eq:corchete entre Ebetas}
\\ F_{\beta_k}F_{\beta_\ell}- \chi(\beta_k,\beta_\ell) F_{\beta_\ell}F_{\beta_k}&= \sum d_{a_{k+1},\ldots,a_{\ell-1}}
F_{\beta_{\ell-1}}^{a_{\ell-1}}\cdots F_{\beta_{k+1}}^{a_{k+1}} \in \mathfrak{u}^-(V),\label{eq:corchete entre Fbetas}
\end{align}
for some scalars $c_{a_{k+1},\ldots,a_{\ell-1}}, d_{a_{k+1},\ldots,a_{\ell-1}}\in \mathbf{k}$.
\end{proposition}
\begin{proof}
Assume first $k=1$. By Lemma \ref{lem:base de ker pi} there exist $\mathtt{c}_{\mathbf{a}}\in \mathbf{k}$ such that
$$ E_{i_1}E_{\beta_\ell}-\chi(\alpha_{i_1},\beta_\ell)E_{\beta_\ell}E_{i_1}= \sum_{\mathbf{a}: \, a_1=0} \mathtt{c}_{\mathbf{a}} \, \mathbf E^{\mathbf{a}}. $$
Let $V'=\rho_{i_\ell}\dots \rho_{i_1}V$. By applying $T_{i_\ell}^-\dots T_{i_1}^-$ to the left hand side we obtain an element of $U^{\le 0}(V')$ . Then $\mathtt{c}_{\mathbf{a}}=0$ if some $a_j>0$ for $j>\ell$ since
\begin{align*}
T_{i_\ell}^-\dots T_{i_1}^-(E_{\beta_j})&=T_{i_\ell}^-\dots T_{i_{j+1}}^-(K_{i_j}^{-1}F_{j})\in U^{\le 0}(V') & \mbox{if } & j\leq \ell, \\
T_{i_\ell}^-\dots T_{i_1}^-(E_{\beta_j})&=T_{i_{\ell+1}}\cdots T_{i_{j-1}}(E_{i_j})\in U^+(V') & \mbox{if } & j> \ell.
\end{align*}
Finally if $\mathtt{c}_{0,a_2,\ldots,a_\ell,0,\ldots, 0}\neq 0$, then
$\sum_{j=2}^\ell a_j\beta_j=\alpha_{i_1}+\beta_\ell$ since $U^+(V)$ is $\mathbb{Z}^{\theta}$-graded so $a_\ell=0$. Therefore we
get \eqref{eq:corchete entre Ebetas} for $c_{a_{k+1},\ldots,a_{\ell-1}}=\mathtt{c}_{0,a_2,\ldots,a_{\ell-1},0,\ldots, 0}$.
The proof of \eqref{eq:corchete entre Fbetas} is analogous.
\end{proof}
Now we introduce algebra filtrations of $U^+(V)$, $U^{\ge0}(V)$ and $U(V)$ related with the PBW basis.
We order $\mathbb{N}_0^M$, $\mathbb{N}_0^{2M+1}$ lexicographically. In particular $\delta_1<\dots<\delta_M$, where $\{\delta_j\}_{1\le j\le M}$ denotes the canonical basis of $\mathbb{Z}^M$ to avoid confusion with the basis $\{\alpha_i\}_{1\le i\le\theta}$ of $\mathbb{Z}^{\theta}$.
\smallbreak\noindent${\mathsf {(i)}\;}$
In $U^+(V)$ set $U^+(V)(\mathbf{a})$ as the subspace spanned by $\mathbf E^{\mathbf{b}}$, $\mathbf{b}\le\mathbf{a}$. This is vector space filtration of $U^+(V)$ so consider the $\mathbb{N}_0^M$-graded vector space
\begin{align*}
\gr U^+(V)&=\oplus_{\mathbf{a}\in\mathbb{N}_0^M} U^+(V)_{\mathbf{a}}, & U^+(V)_{\mathbf{a}}&=U^+(V)(\mathbf{a})/\sum_{\mathbf{b}<\mathbf{a}} U^+(V)(\mathbf{b}).
\end{align*}
\smallbreak\noindent${\mathsf {(ii)}\;}$
For $U^{\ge0}(V)$, $U^{\ge0}(V)(\mathbf{a})$ is the subspace spanned by $\mathbf E^{\mathbf{b}}K^\alpha$, $\mathbf{b}\le\mathbf{a}$, $\alpha\in\mathbb{Z}^{\theta}$. In particular $U^{\ge0}(V)(0)=U^{+0}(V)$.
\smallbreak\noindent${\mathsf {(iii)}\;}$
For each $\mathbf E^{\mathbf{a}}K^\alpha L^\beta\mathbf F^{\mathbf{b}}\in U(V)$ set
$$ d(\mathbf E^{\mathbf{a}}K^\alpha L^\beta\mathbf F^{\mathbf{b}})=\left(\sum_{j=1}^M (a_j+b_j)\hgt(\beta_j),a_1,\dots,a_M,b_1,\dots,b_M \right) \in\mathbb{N}_0^{2M+1}. $$
For each $\mathbf{u}\in\mathbb{N}_0^{2M+1}$ let $U(V)(\mathbf{u})$ be the subspace spanned by $\mathbf E^{\mathbf{a}}K^\alpha L^\beta\mathbf F^{\mathbf{b}}$, $d(\mathbf E^{\mathbf{a}}K^\alpha L^\beta\mathbf F^{\mathbf{b}})\le\mathbf{u}$. Then take the associated $\mathbb{N}_0^{2M+1}$-graded vector space
\begin{align*}
\gr U(V)&=\oplus_{\mathbf{u}\in\mathbb{N}_0^{2M+1}} U(V)_{\mathbf{u}}, & U(V)_{\mathbf{u}}&=U(V)(\mathbf{u})/\sum_{\mathbf{v}<\mathbf{u}} U(V)(\mathbf{v}).
\end{align*}
Next result generalizes \cite[Proposition 1.7]{DK-root of 1}, \cite[Proposition 10.1]{DP-quantum}
\begin{proposition}\label{prop:filtracion via PBW}
The $\mathbb{N}_0^M$-filtrations on $U^{+}(V)$, $U^{\ge0}(V)$ and the $\mathbb{N}_0^{2M+1}$-filtration on $U(V)$ are algebra filtrations.
\end{proposition}
\begin{proof}
First we consider $U^{+}(V)$. We claim that $\mathbf E^{\mathbf{a}}\mathbf E^{\mathbf{b}}\in U^+(V)(\mathbf{a}+\mathbf{b})$. Let $\mathbf{a}=\delta_k$, $\mathbf{b}=\delta_\ell$, $1\le k,\ell \le M$.
Note that $U^+(V)(\delta_k)= \mathbf{k} E_{\beta_k}$, and $E_{\beta_k}E_{\beta_\ell}\in U^+(V)(\delta_k+\delta_\ell)$ if $k\geq \ell$ by definition. Also $E_{\beta_k}E_{\beta_\ell}\in U^+(V)(\delta_k+\delta_\ell)$ if $k<\ell$ by Proposition \ref{prop: corchete entre Ebetas}, since
\begin{align*}
\sum_{j=k+1}^{\ell+1} a_j\delta_j&<\delta_k+\delta_\ell & \mbox{ so }& E_{\beta_{\ell-1}}^{a_{\ell-1}} \cdots E_{\beta_{k+1}}^{a_{k+1}}\in \sum_{\mathbf{v}<\delta_k+\delta_\ell} U^+(V)(\mathbf{v}).
\end{align*}
Using this case we can reorder the PBW generators of $\mathbf E^{\mathbf{a}}\mathbf E^{\mathbf{b}}$ for any $\mathbf{a},\mathbf{b}$.
The proof for $U^{\ge0}(V)$ follows from the previous case since $K_i$ $q$-commutes with all the elements of $U^+(V)$. For $U(V)$ we argue as above and reduce the proof to the product between $E_{\beta_k}$, $F_{\beta_\ell}$. By Corollary \ref{coro:FE=EF+cosas},
\begin{align}\label{eq:Fbeta Egamma}
F_{\beta_\ell}E_{\beta_k}\in E_{\beta_k}F_{\beta_\ell}+ \sum_{\mathbf{v}: v_1\leq \hgt(\beta_\ell)+\hgt(\beta_k)-2}U(V)(\mathbf{v})
\end{align}
so
$ F_{\beta_\ell}E_{\beta_k}\in U(V)(\hgt(\beta_\ell)+\hgt(\beta_k),\delta_k,\delta_\ell)$.
\end{proof}
We give a presentation of the corresponding graded algebras.
\begin{corollary}\label{coro:graduada via PBW, presentacion}
${\mathsf {(i)}\;}$ The algebra $\gr U^+(V)$ is presented by generators $\mathtt E_k$, $1\le k\le M$ and relations
\begin{align}\label{eq:rels grU+}
\mathtt E_k \mathtt E_\ell&=\chi(\beta_k,\beta_\ell)\, \mathtt E_\ell \mathtt E_k, \quad k<\ell, & \mathtt E_k^{N_k}&=0,\quad \beta_k\notin\mathcal{O}(V).
\end{align}
\smallbreak
\noindent${\mathsf {(ii)}\;}$ The algebra $\gr U^{\ge 0}(V)$ is presented by generators $\mathtt E_k$, $1\le k\le M$, $K_i^{\pm}$, $i\in\I$, and relations \eqref{eq:rels grU+},
\begin{align}\label{eq:rels grU ge0 - 1}
K_i K_j &= K_j K_i, & K_iK_i^{-1}&=K_i^{-1}K_i=1, \\
K_i\mathtt E_k &=\chi(\alpha_i,\beta_k)\, \mathtt E_k K_i, & 1\le k\le & M, \, i,j\in\I. \label{eq:rels grU ge0 - 2}
\end{align}
\end{corollary}
\begin{proof}
For ${\mathsf {(i)}\;}$ Let $\mathcal{F}$ be the free algebra generated by $\mathtt E_k$, $1\le k\le M$ and $\pi:\mathcal{F}\to \gr U^+(V)$ the algebra map such that $\pi(\mathtt E_k)=E_{\beta_k}$. By \eqref{eq:corchete entre Ebetas}, $E_{\beta_k}E_{\beta_\ell}= \chi(\beta_k,\beta_\ell)\, E_{\beta_\ell}E_{\beta_k}$ holds in $\gr U^+(V)$, and also $E_{\beta_k}^{N_k}=0$ for $\beta_k\notin\mathcal{O}(V)$ by Remark \ref{rem:PRV=0 si alpha no cartan}. Then $\pi$ factors through the algebra defined by the relations \eqref{eq:rels grU+}, which is the quotient of a $q$-polynomial algebra of $M$ generators by $\mathtt E_k^{N_k}$ for those generators corresponding to $\beta_k\notin\mathcal{O}(V)$, so it has a basis
$$ \{\mathtt E_M^{a_m}\dots \mathtt E_1^{a_1} | a_k\in\mathbb{N}_0, a_k<N_k \mbox{ if }\beta_k\notin\mathcal{O}(V)\}, $$
As $\mathbf E^{\mathbf{a}}\in U^+(V)(\mathbf{a})-\sum_{\mathbf{b}<\mathbf{a}} U^+(V)(\mathbf{b})$, $\gr U^+(V)$ has a basis corresponding to the images of \eqref{eq:base de U+i}. But
$\pi(\mathtt E_M^{a_m}\dots \mathtt E_1^{a_1})=\mathbf E^{\mathbf{a}}$ so $\pi$ is an isomorphism. The proof of ${\mathsf {(ii)}\;}$ follows similarly.
\end{proof}
\begin{corollary}\label{coro:graduada via PBW, presentacion Uchi}
The algebra $\gr U(V)$ is presented by generators $\mathtt E_k$, $\mathtt F_k$, $1\le k\le M$, $K_i$, $K_i^{-1}$, $L_i$, $L_i^{-1}$, $i\in\I$, and relations
\begin{align*}
XY&=YX, & X,Y \in & \{ K_i^{\pm1}, L_i^{\pm1}: 1 \leq i \leq \theta \}, \\
K_iK_i^{-1}&=L_iL_i^{-1}=1, & \mathtt E_k\mathtt F_\ell&=\mathtt F_\ell\mathtt E_k,
\\ K_i\mathtt E_k&=\chi(\alpha_i,\beta_k)\, \mathtt E_k K_i, & L_i \mathtt E_k&=\chi(\beta_k,-\alpha_i)\,\mathtt E_k L_i,
\\ K_i\mathtt F_k&=\chi(-\alpha_i,\beta_k)\,\mathtt F_kK_i, & L_i\mathtt F_k&=\chi(\beta_k,\alpha_i)\,\mathtt F_kL_i,
\\ \mathtt E_k \mathtt E_\ell&=\chi(\beta_k,\beta_\ell)\, \mathtt E_\ell \mathtt E_k, \quad k<\ell, & \mathtt E_k^{N_k}&=0,\quad \beta_k\notin\mathcal{O}(V),
\\ \mathtt F_k \mathtt F_\ell&=\chi(\beta_\ell,\beta_k)\, \mathtt F_\ell \mathtt F_k, \quad k<\ell, & \mathtt F_k^{N_k}&=0,\quad \beta_k\notin\mathcal{O}(V).
\end{align*}
\end{corollary}
\begin{proof} The proof is analogous to the previous case if we check that $E_{\beta_k} F_{\beta_\ell}= F_{\beta_\ell}E_{\beta_k}$ in $\gr U(V)$; this relation follows by \eqref{eq:Fbeta Egamma}.
\end{proof}
\medskip
\begin{theorem}\label{thm:noetheriano}
The algebras $U^{+}(V)$, $U^{\ge0}(V)$, $U(V)$ are Noetherian.
\end{theorem}
\begin{proof}
It suffices to prove that the graded algebras are Noetherian. As
\begin{itemize}
\item $\gr U^{+}(V)$ is the quotient of a quantum affine space on $M$ generators by powers of some generators,
\item $\gr U^{\ge0}(V)$ is the localization (we add the inverses of the $\theta$ generators $K_i$) of the quotient of a quantum affine space on $M+\theta$ generators by powers of some generators, and
\item $\gr U(V)$ is the localization of the quotient of a quantum affine space on $2M+2\theta$ generators by powers of some generators,
\end{itemize}
the three algebras $\gr U^{+}(V)$, $\gr U^{\ge0}(V)$, $\gr U(V)$ are Noetherian.
\end{proof}
Now we compute the Gelfand-Kirillov dimension of these algebras. We refer to \cite{KL} for the definition and properties.
\begin{theorem}\label{thm:GKdim finita}
The Gelfand-Kirillov dimension of $U^{+}(V)$, $U^{\ge0}(V)$, $U(V)$ are, respectively, $|\mathcal{O}(V)|$, $|\mathcal{O}(V)|+\theta$, $2|\mathcal{O}(V)|+2\theta$.
\end{theorem}
\begin{proof}
By \cite[Proposition 6.6]{KL}, $\GKdim U^{+}(V)=\GKdim \gr U^{+}(V)$.
The subalgebra $S^+(V)$ of $\gr U^{+}(V)$ generated by $\mathtt E_k$, with $\beta_k\in\mathcal{O}(V)$, is a quantum affine space in
$|\mathcal{O}(V)|$ generators and $\gr U^{+}(V)$ is a free $S^+(V)$-module of rank $\prod\limits_{k:\beta_k\notin \mathcal{O}(V)} N_k$, so $ \GKdim \gr U^{+}(V)=\GKdim S^+(V)= |\mathcal{O}(V)|$.
For $U^{\ge 0}(V)$, let $S^{\ge0}(V)$ be the subalgebra of $\gr U^{\ge0}(V)$ generated by $K_i$, $i\in\I$, $\mathtt E_k$ if $\beta_k\in\mathcal{O}(V)$; it is the localization of a quantum affine space with $|\mathcal{O}(V)|+\theta$ generators, and $\gr U^{\ge 0}(V)$ is a free $S^{\ge 0}(V)$-module of rank $\prod\limits_{k:\beta_k\notin \mathcal{O}(V)} N_k$, so $ \GKdim U^{\ge 0}(V)=\GKdim \gr U^{\ge 0}(V)=\GKdim S^{\ge 0}(V)= |\mathcal{O}(V)|+\theta$.
The proof for $U(V)$ follows analogously.
\end{proof}
\section{Power root vectors on distinguished pre-Nichols algebras}\label{section:subalgebra Z}
Now we study the subalgebra $Z(V)$ of $U(V)$ generated by $E_\alpha^{N_\alpha}$, $F_\alpha^{N_\alpha}$, $K_\alpha^{N_\alpha}$, $L_\alpha^{N_\alpha}$, $\alpha\in\mathcal{O}(V)$. First we describe the product of these elements. Then we give a general formula for the composition of the coproduct with $T_i^{V}$ on the whole $U^+_{-i}(V)$, and finally apply this formula to show that $Z(V)$ is a Hopf subalgebra. In what follows
\begin{itemize}
\item $Z^+(V)$ and $Z^-(V)$ are the subalgebras generated by $E_\alpha^{N_\alpha}$, respectively $F_\alpha^{N_\alpha}$, $\alpha\in\mathcal{O}(V)$.
\item $\Gamma(V)$ is the subgroup of $\mathbb{Z}^{2\theta}$ generated by $K_\alpha^{N_\alpha}$, $L_\alpha^{N_\alpha}$, $\alpha\in\mathcal{O}(V)$.
\item $S(V)$ is a set of representatives of $\mathbb{Z}^{2\theta}/\Gamma(V)$.
\end{itemize}
The multiplication gives an isomorphism of $\mathbb{Z}^{\theta}$-graded vector spaces $Z^+(V)\otimes \mathbf{k} \Gamma(V)\otimes Z^-(V)\simeq Z(V)$.
\subsection{Algebra structure of $Z(V)$}
Power root vectors are central elements of quantized enveloping algebras $U_q(\mathfrak g)$, $q$ a root of unity \cite[Chapter 19]{DP-quantum}. In the general case they $q$-commute with the elements of $U(V)$.
\begin{proposition}\label{prop:power roots q-commute}
Let $\beta\in\mathcal{O}(V)$.
\noindent${\mathsf {(i)}\;}$ For each $X\in U^+(V)$ homogeneous of degree $\gamma$,
\begin{align*}
E_\beta^{N_\beta} X&= \chi(N_\beta\beta,\gamma)\, XE_\beta^{N_\beta}, & F_\beta^{N_\beta} X&= XF_\beta^{N_\beta}.
\end{align*}
\noindent${\mathsf {(ii)}\;}$ For each $Y\in U^-(V)$ homogeneous of degree $\gamma$,
\begin{align*}
E_\beta^{N_\beta} Y&= YE_\beta^{N_\beta}, & F_\beta^{N_\beta} Y&= \chi(\gamma,N_\beta\beta)\, YF_\beta^{N_\beta}.
\end{align*}
\end{proposition}
\begin{proof}
If $\beta=\alpha_i$, then $E_i^{N_i} F_j=F_jE_i^{N_i}$ for $j\neq i$, and $E_i^{N_i} F_i=F_iE_i^{N_i}$ by \eqref{eq:simple power root contra Fp}, since $(N_i)_{q_{ii}}=0$. Therefore $E_i^{N_i} Y= YE_i^{N_i}$ for all $Y\in U^-(V)$.
For each $j\neq i$, $0=(\ad_c E_i)^{N_i} E_j= E_i^{N_i} E_j-q_{ij}^{N_i} E_j E_i^{N_i}$ by \eqref{eq:expresion adjunta}, since $N_i\geq 1-a_{ij}$.
Then $E_i^{N_i} X= \chi(N_i\alpha_i,\gamma) XE_i^{N_i}$ for all $X\in U^+(V)_\gamma$, $\gamma\in\mathbb{N}_0$.
If $\beta=s_{i_1}\cdots s_{i_{k-1}}(\alpha_{i_k})$, then $E_\beta=T_{i_1}\cdots T_{i_{k-1}}(\underline E_{i_k})$. Let $v=s_{i_{k-1}}\dots s_{i_1}$. If $j$ is such that $v(\alpha_j)\in\mathbb{N}_0^\theta$, then $T_{i_{k-1}}^-\dots T_{i_1}^-(E_j)\in U^+(V')_{v(\alpha_j)}$ by \cite[Proposition 6.15]{H-isom}, so we can use the previous case:
\begin{align*}
E_\beta^{N_\beta} E_j &= T_{i_1}\cdots T_{i_{k-1}}\left( v^* \chi(N_{i_k}\alpha_{i_k},v(\alpha_j)) \underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_1}^-(E_j) \right) \\
&= \chi(N_\beta\beta,\alpha_j) E_j E_\beta^{N_\beta}.
\end{align*}
As also $T_{i_{k-1}}^-\dots T_{i_1}^-(F_j)\in U^-(V')_{-v(\alpha_j)}$, we have that
\begin{align*}
E_\beta^{N_\beta} F_j &= T_{i_1}\cdots T_{i_{k-1}}\left(\underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_1}^-(F_j) \right) \\
&= T_{i_1}\cdots T_{i_{k-1}}\left(T_{i_{k-1}}^-\dots T_{i_1}^-(F_j) \underline E_{i_k}^{N_{i_k}}\right) = F_j E_\beta^{N_\beta}.
\end{align*}
If $j$ satisfies $v(\alpha_j)\in-\mathbb{N}_0^\theta$, then $s_{i_{t-1}}\cdots s_{i_1}(\alpha_j)\in\mathbb{N}_0^\theta$, $s_{i_t}\cdots s_{i_1}(\alpha_j)\in-\mathbb{N}_0^\theta$ for $t\geq 0$ minimal. Therefore $s_{i_{t-1}}\cdots s_{i_1}(\alpha_j)=\alpha_{i_t}$, so $T_{i_{t-1}}^-\dots T_{i_1}^-(E_j)=c\,E_{i_t}$ for some $c\in \mathbf{k}^\times$, and
$$ T_{i_{k-1}}^-\dots T_{i_1}^-(E_j)= T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(c\, K_t^{-1}F_t) \in T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(K_t^{-1})U^-(V'), $$
because $s_{i_{k-1}}\dots s_{i_{t+1}}(-\alpha_t)=w(\alpha_j)\in-\mathbb{N}_0^\theta$, so
\begin{align*}
E_\beta^{N_\beta} & E_j = T_{i_1}\cdots T_{i_{k-1}}\left(\underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_1}^-(E_j) \right) \\
&= c\, T_{i_1}\cdots T_{i_{k-1}}\left(\underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_t}^-(cE_t) \right) \\
&= c\, T_{i_1}\cdots T_{i_{k-1}}\left(\underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(K_t^{-1})T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(F_t) \right) \\
&= c\, T_{i_1}\cdots T_{i_{k-1}}\big( w^* \chi(N_{i_k}\alpha_{i_k},w(\alpha_j)) T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(K_t^{-1}) \\
& \qquad \underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(F_t) \big) \\
&= c\, T_{i_1}\cdots T_{i_{k-1}}\big( w^* \chi(N_{i_k}\alpha_{i_k},w(\alpha_j)) T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(K_t^{-1}) \\
& \qquad T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(F_t) \underline E_{i_k}^{N_{i_k}} \big) = \chi(N_\beta\beta,\alpha_j) E_j E_\beta^{N_\beta}.
\end{align*}
Analogously, $T_{i_{t-1}}^-\dots T_{i_1}^-(E_j)=c'\,E_{i_t}$ for some $c'\in \mathbf{k}^\times$, so
$$ T_{i_{k-1}}^-\dots T_{i_1}^-(F_j)= T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(c'\, E_t L_t^{-1}) \in U^+(V')T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(L_t^{-1}), $$
because $s_{i_{k-1}}\dots s_{i_{t+1}}(-\alpha_t)=w(\alpha_j)\in-\mathbb{N}_0^\theta$, and
\begin{align*}
E_\beta^{N_\beta} & E_j = T_{i_1}\cdots T_{i_{k-1}}\left(\underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_1}^-(E_j) \right) \\
& = T_{i_1}\cdots T_{i_{k-1}}\left(\underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_t}^-(cE_t) \right) \\
&= c'\, T_{i_1}\cdots T_{i_{k-1}}\left(\underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(E_t)T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(L_t^{-1}) \right) \\
&= c'\, T_{i_1}\cdots T_{i_{k-1}}\Big( w^* \chi(N_{i_k}\alpha_{i_k},w(\alpha_j)) T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(E_t) \\
&\qquad \underline E_{i_k}^{N_{i_k}} T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(L_t^{-1}) \Big) \\
&= c'\, T_{i_1}\cdots T_{i_{k-1}}\left( T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(E_t) T_{i_{k-1}}^-\dots T_{i_{t+1}}^-(L_t^{-1}) \underline E_{i_k}^{N_{i_k}} \right) \\
&= \chi(N_\beta\beta,\alpha_j) E_j E_\beta^{N_\beta},
\end{align*}
which completes the proof for $E_\beta^{N_\beta}$. The proof for $F_\beta^{N_\beta}$ is analogous.
\end{proof}
\begin{corollary}\label{coro:derivadas PRV}
For all $\alpha\in\mathcal{O}(V)$ and all $j\in\I$, $ \partial_j^K(E_\alpha^{N_\alpha})=\partial_j^L(E_\alpha^{N_\alpha})=0.$
\end{corollary}
\begin{proof} Let $i=i_1$.
The proof for $\alpha=\alpha_{i}$ is direct. If $\alpha\neq\alpha_i$, then set $\beta=s_i(\alpha)\in\mathcal{O}(\rho_i(V))$, so $E_\alpha=T_i(E_\beta)$, $N_\alpha=N_\beta$. By Proposition \ref{prop:power roots q-commute},
\begin{align*}
\partial_i^K(E_\alpha^{N_\alpha})K_i- &L_i^{-1}\partial_i^L(E_\alpha^{N_\alpha})= E_\alpha^{N_\alpha} F_i-F_iE_\alpha^{N_\alpha} \\
&= T_i\left( E_\beta^{N_\beta} E_iL_i^{-1} - E_iL_i^{-1}E_\beta^{N_\beta} \right) \\
&= T_i\left( s_i^*\chi( N_\beta\beta,\alpha_i) E_iE_\beta^{N_\beta} L_i^{-1} - E_iL_i^{-1}E_\beta^{N_\beta} \right)=0.
\end{align*}
Then $\partial_i^K(E_\alpha^{N_\alpha})=\partial_i^L(E_\alpha^{N_\alpha})=0$. For $j\neq i$,
\begin{align*}
\partial_j^K(E_\alpha^{N_\alpha})K_j- &L_j^{-1}\partial_j^L(E_\alpha^{N_\alpha})= E_\alpha^{N_\alpha} F_j-F_jE_\alpha^{N_\alpha} \\
&= T_{i}\left( E_\beta^{N_\beta} F_{i}^{-a_{ij}^{V}} F_j - F_{i}^{-a_{ij}^{V}} F_jE_\beta^{N_\beta} \right)=0,
\end{align*}
so $\partial_j^K(E_\alpha^{N_\alpha})=\partial_j^L(E_\alpha^{N_\alpha})=0$.
\end{proof}
We deduce the freeness of $U(V)$ as $Z(V)$-module.
\begin{theorem}\label{thm:root vectors almost central, freeness}
The set
\begin{align*}
& \left\{ \left(\prod_{\beta\in\mathcal{O}(V)}E_{\beta}^{k_\beta N_\beta}\right) \gamma \left(\prod_{\beta\in\mathcal{O}(V)}F_{\beta}^{l_\beta N_\beta}\right): k_\beta,l_\beta\in\mathbb{N}_0,\gamma\in\Gamma(V) \right\},
\end{align*}
is a basis of $Z(V)$, where the order on the set $\mathcal{O}(V)$ is arbitrary.
Moreover $U(V)$ is a free $Z(V)$-module with basis:
\begin{align*}
& \left\{ E_{\beta_M}^{a_M}E_{\beta_{M-1}}^{a_{M-1}} \cdots E_{\beta_1}^{a_1} s F_{\beta_M}^{b_M}F_{\beta_{M-1}}^{b_{M-1}} \cdots
F_{\beta_1}^{b_1}\, | 0\leq a_k,b_k < N_k , s\in S(V) \right\}.
\end{align*}
\end{theorem}
\begin{proof}
By Proposition \ref{prop:power roots q-commute}, the generators of the subalgebra $Z(V)$ $q$-commute with all the elements of $U(V)$.
Then apply Theorem \ref{thm: PBW bases preNichols}.
\end{proof}
\begin{remark}
The algebra $Z(V)$ is not necessarily central since we can have $\chi(N_\beta\beta,\alpha_j)\neq1$ for some $\beta\in\mathcal{O}(V)$, $1\le j\le \theta$.
\end{remark}
\begin{lemma}\label{lema:caracter sim trivial en Nbeta beta}
Let $\beta\in\mathcal{O}(V)$. For every $\alpha\in\mathbb{Z}^\theta$, $\chi(N_\beta\beta,\alpha)\chi(\alpha,N_\beta\beta)=1$.
\end{lemma}
\begin{proof}
If $\beta=\alpha_i$, then $\chi(N_i\alpha_i,\alpha_i)\chi(\alpha_i,N_i\alpha_i)=q_{ii}^{2 N_i}=1$ and for $j\neq i$,
$$\chi(N_i\alpha_i,\alpha_j)\chi(\alpha_j,N_i\alpha_i)=(q_{ij}q_{ji})^{N_i}=q_{ii}^{c_{ij}^{V} N_i}=1,$$
so $\chi(N_i\alpha_i,\alpha)\chi(\alpha,N_i\alpha_i)=1$ for every $\alpha\in\mathbb{Z}^\theta$.
If $\beta\neq\alpha_i$, then $\beta=w(\alpha_i)$ for some $w\in\Hom(\mathcal{W},V)$ and $\alpha_i\in\mathcal{O}(w^{-1}(V))$, so $N_\beta=N_i$. Therefore
\begin{align*}
\chi(N_\beta\beta,\alpha)\chi & (\alpha,N_\beta\beta)=\chi(N_i w(\alpha_i),\alpha)\chi(\alpha,N_i w(\alpha_i))\\
&=(w^{-1})^*\chi(N_i \alpha_i,w^{-1}(\alpha)) (w^{-1})^*\chi(w^{-1}(\alpha),N_i \alpha_i)=1,
\end{align*}
for all $\alpha\in\mathbb{Z}^\theta$, by applying the previous step to $\alpha_i\in\mathcal{O}(w^{-1}(V))$.
\end{proof}
\subsection{On the coproduct of $U(V)$}
We want to factorize the composition $\underline{\Delta}\circ T_i: U_{-i}^+(V)\to U^+(V)\otimes U_{+i}^+(V)$ as in \cite{HS-coideal subalg}. Similar formula was introduced in \cite[Proposition 5.3.4]{L-libro} for quantized enveloping algebras. Such factorization is interpreted as equivalences between the corresponding categories of Yetter-Drinfeld modules \cite{BLS,HS-general}. One of the factors is $T_i\otimes T_i$ but viewed as an algebra map between \emph{braided} structures.
\begin{itemize}
\item $U_{-i}^+(V)\underline{\otimes} U_{-i}^+(V)$ denotes the tensor product $U_{-i}^+(V)\otimes U_{-i}^+(V)$ with the algebra structure in $^{\mathcal{B}_{i}^{\bop}\# \mathbf{k}\mathbb{Z}^\theta}_{\mathcal{B}_{i}^{\bop}\# \mathbf{k}\mathbb{Z}^\theta}\mathcal{YD}$, see Remark \ref{rem:algebra de Hopf YD para U-i}, and
\item $U_{+i}^+(V)\underline{\otimes} U_{+i}^+(V)$ denotes the tensor product $U_{+i}^+(V)\otimes U_{+i}^+(V)$ with the algebra structure in $^{\mathcal{B}_{i}\# \mathbf{k}\mathbb{Z}^\theta}_{\mathcal{B}_{i}\# \mathbf{k}\mathbb{Z}^\theta}\mathcal{YD}$, see Remark \ref{rem:algebra de Hopf YD para U+i}.
\end{itemize}
\begin{lemma}\label{lemma: Ti ot Ti es de algebras}
$T_i\otimes T_i: U_{-i}^+(V)\underline{\otimes} U_{-i}^+(V)\to U_{+i}^+(V)\underline{\otimes} U_{+i}^+(V)$ is an algebra map.
\end{lemma}
\begin{proof}
Let $x,y,w,z\in U_{-i}^+(V)$, where $y$, $w$ are homogeneous of degrees $\beta,\gamma\in\mathbb{N}_0^\theta$, respectively. As $(\ad_c E_i)^{N_i}\equiv 0$,
\begin{align*}
T_i & \otimes T_i (x\otimes y) \cdot \, T_i\otimes T_i(w\otimes z)= T_i(x) \left( T_i(y)_{(-1)}\rightharpoonup T_i(w) \right) \underline{\otimes} T_i(y)_{(0)} T_i(z) \\
& \overset{\eqref{eq:accion-coaccion a izquierda}}= \sum_{n=0}^{N_i-1} s_i^*\chi(s_i(\beta)-n\alpha_i,s_i(\gamma)) \, T_i(x) (\ad_{\underline c} \underline E_i)^n T_i(w)\underline{\otimes} \, (T_i(y) \triangleleft \mathfrak{E}_i^{(n)}) T_i(z) \\
& \overset{\eqref{eq:producto coproducto dual},\eqref{eq:Ti con delta K}}= \sum_{n=0}^{N_i-1} \frac{(-1)^nq_{ii}^{\frac{n(n-1)}{2}}\chi(\beta+n\alpha_i,\gamma)}{(n)_{q_{ii}}! \chi(\alpha_i,\gamma)^{n}}
\, T_i(x) T_i((\partial_i^K)^n w) \\
& \qquad \qquad \qquad \underline{\otimes} \left((\partial_i^L)^n T_i(y) \right) T_i(z) \\
& \overset{\eqref{eq:producto coproducto dual}}= \sum_{n=0}^{N_i-1} (-1)^nq_{ii}^{\frac{n(n-1)}{2}}\chi(\beta,\gamma)
\, T_i(x) T_i(\mathfrak{E}_i^{(n)} \triangleright w)\underline{\otimes} \left((\partial_i^L)^n T_i(y) \right) T_i(z) \\
& \overset{\eqref{eq:Ti con delta L}}= \sum_{n=0}^{N_i-1} \frac{\chi(\beta,\gamma)}{\chi(\beta,\alpha_i)^{n}}
\, T_i(x) T_i(\mathfrak{E}_i^{(n)} \triangleright w)\underline{\otimes} T_i \left(E_i^n \rightharpoonup ' y \right) T_i(z) \\
& = \sum_{n=0}^{N_i-1} \chi(\beta,\gamma-n\alpha_i)
\, T_i \left( x \left(\mathfrak{E}_i^{(n)} \triangleright w \right) \right)\underline{\otimes} T_i \left( \left(E_i^n \rightharpoonup ' y \right) z \right) \\
& = T_i\otimes T_i \left( (x\otimes y) \cdot (w\otimes z) \right),
\end{align*}
so the proof is complete.
\end{proof}
Let $Y\in U^+_{+i}(V)_\beta$, $\beta\in\mathbb{N}_0^\theta$. As $Y \triangleleft\mathfrak{E}_i^{(n)}\in U(V)_{\beta-n\alpha_i}$, there exists $k\in\mathbb{N}$ such that $Y \triangleleft\mathfrak{E}_i^{(n)} =0$ for all $n\geq k$. Then there exists a well-defined map $\mathfrak{R}_i:U_{+i}^+(V)\otimes U_{+i}^+(V)\to U^+(V)\otimes U_{+i}^+(V)$ such that
\begin{align}\label{eq:def Ri}
\mathfrak{R}_i^{V}(x\otimes y)&= \sum_{k\geq0} x E_i^k\otimes y \triangleleft\mathfrak{E}_i^{(k)}, & x,y &\in U^+_{+i}(V).
\end{align}
Compare with \cite[(3.4)]{HS-coideal subalg}.
\begin{lemma}\label{lemma:Rfi es de algebras}
$\mathfrak{R}_i^{V}:U_{+i}^+(V)\underline{\otimes} U_{+i}^+(V) \to U^+(V)\otimes U_{+i}^+(V)$ is an algebra map.
\end{lemma}
Here $U^+(V)\otimes U_{+i}^+(V)$ is a subalgebra of $U^+(V)\otimes U^+(V)$, where $U^+(V)\otimes U^+(V)$ has the algebra structure such that
$ (x\otimes y)(x'\otimes y)= \chi(\beta,\gamma) xx'\otimes yy$
for $x,x',y,y'\in U^+(V)$, $y$, $x'$ homogeneous of degree $\beta$, $\gamma$, respectively.
\begin{proof}
Set $x,y,w,z\in U_{+i}^+(V)$, $y$, $w$ homogeneous of degrees $\beta,\gamma\in\mathbb{N}_0^\theta$. Then
\begin{align}\label{eq:formula adjunta}
E_i^n \, w
&= \sum_{r=0}^n \binom{n}{r}_{q_{ii}}\chi(\alpha_i,\gamma)^{n-r} (\ad_c E_i)^r w \, E_i^{n-r}.
\end{align}
for all $n\in\mathbb{N}$, so
\begin{align*}
\mathfrak{R}_i^{V} &(x\otimes y) \cdot \mathfrak{R}_i^{V}(w\otimes z)= \\
&= \sum_{j,k\geq 0} \chi(\beta-j\alpha_i,\gamma+k\alpha_i) \, xE_i^j w E_i^k \otimes (y\triangleleft\mathfrak{E}_i^{(j)}) (z\triangleleft\mathfrak{E}_i^{(k)}) \\
& \overset{\eqref{eq:formula adjunta}} = \sum_{j,k\geq 0}
\, x \left( \sum_{r=0}^{\min \{N_i-1,j\} } \binom{j}{r}_{q_{ii}}
\frac{\chi(\beta,\gamma)\chi(\beta,\alpha_i)^k}{\chi(\alpha_i,\gamma)^rq_{ii}^{jk} }
(\ad_c E_i)^r w E_i^{k+j-r} \right)
\\ &\qquad \qquad\qquad \otimes (y\triangleleft\mathfrak{E}_i^{(j)}) (z\triangleleft\mathfrak{E}_i^{(k)})
\\
& \overset{\eqref{eq:producto coproducto dual}} = \sum_{j,k\geq 0} \sum_{r=0}^{\min \{N_i-1,j\} }
\frac{\chi(\beta,\gamma+k\alpha_i)}{\chi(\alpha_i,\gamma)^r q_{ii}^{jk} }
x(\ad_c E_i)^r w E_i^{k+j-r} \\
&\qquad \qquad\qquad \otimes ((y\triangleleft\mathfrak{E}_i^{(r)}) \triangleleft\mathfrak{E}_i^{(j-r)}) (z\triangleleft\mathfrak{E}_i^{(k)}).
\end{align*}
On the other hand,
\begin{align*}
\mathfrak{R}_i^{V} &\left( (x\otimes y) \cdot (w\otimes z)\right) =\\
&\overset{\eqref{eq:accion-coaccion a izquierda}}= \mathfrak{R}_i^{V} \left( \sum_{r=0}^{N_i} \chi(\beta-r\alpha_i,\gamma) \, x(\ad_c E_i)^r w \otimes (y \triangleleft \mathfrak{E}_i^{(r)})z \right) \\
& \overset{\eqref{eq:def Ri}} = \sum_{r=0}^{N_i}\sum_{t\geq 0} \chi(\beta-r\alpha_i,\gamma) \, x(\ad_c E_i)^r w E_i^t \otimes \left((y \triangleleft \mathfrak{E}_i^{(r)})z \right)\triangleleft \mathfrak{E}_i^{(t)}
\\ & \overset{\eqref{eq:Efi como skew derivation}} = \sum_{r=0}^{N_i}\sum_{k,l\geq 0} \chi(\beta-r\alpha_i,\gamma)\chi(\beta-(r+l)\alpha_i,k\alpha_i) \, x(\ad_c E_i)^r w E_i^{k+l}
\\ & \qquad \qquad \qquad \otimes ((y \triangleleft \mathfrak{E}_i^{(r)}) \triangleleft \mathfrak{E}_i^{(l)} )z \triangleleft \mathfrak{E}_i^{(k)}
\\ & = \sum_{r=0}^{N_i}\sum_{k,l\geq 0} \frac{\chi(\beta,\gamma+k\alpha_i)}{\chi(\alpha_i,\gamma)^r q_{ii}^{(r+l)k}} \, x(\ad_c E_i)^r w E_i^{k+l} \otimes ((y \triangleleft \mathfrak{E}_i^{(r)}) \triangleleft \mathfrak{E}_i^{(l)} )z \triangleleft \mathfrak{E}_i^{(k)},
\end{align*}
so $\mathfrak{R}_i^{V} (x\otimes y) \cdot \mathfrak{R}_i^{V}(w\otimes z)=\mathfrak{R}_i^{V} \left( (x\otimes y) \cdot (w\otimes z)\right)$.
\end{proof}
Recall the map $\Delta_i:U_{+i}^+(V) \to U_{+i}^+(V)\underline{\otimes} U_{+i}^+(V)$ introduced in Remark \ref{rem:algebra de Hopf YD para U-i}. Now we prove the factorization of $\underline{\Delta}\circ T_i$.
\begin{theorem}\label{thm:factorization delta Ti}
For every $E\in U_{-i}^+(V)$,
\begin{equation}\label{eq:factorization delta Ti}
\underline{\Delta}\circ T_i (E)= \mathfrak{R}_i^{\rho_i(V)} \circ (T_i\otimes T_i)\circ \Delta_i(E).
\end{equation}
\end{theorem}
\begin{proof}
It is enough to prove the formula for $E=E_{j,n}^-$, $j\neq i$, $0\leq n\leq -c_{ij}^{V}$, since they generate $U_{-i}^+(V)$ as an algebra and both $\underline{\Delta}\circ T_i$, $\mathfrak{R}_i^{\rho_i(V)} \circ (T_i\otimes T_i)\circ \Delta_i$ are algebra maps by Lemmas \ref{lemma: Ti ot Ti es de algebras} and \ref{lemma:Rfi es de algebras}. Let
$$ \kappa_{i,j;n}:= \underline{q}_{ii}^{\,\, n} \left(\prod_{t=-c_{ij}^{V}-n}^{-c_{ij}^{V}-1} (t+1)_{\underline{q}_{ii}}(1-\underline{q}_{ii}^{\,\, t}\underline{q}_{ij}\underline{q}_{ji}) \right).$$
By \eqref{eq:Ti(E-)=E+} and \eqref{eq:coproducto Ei+},
\begin{align*}
\kappa_{i,j;n}^{-1} & \underline{\Delta}\circ T_i(E_{j,n}^-) = \underline{\Delta}(\underline E_{j,-c_{ij}^{V}-n}^+) = \underline E_{j,-c_{ij}^{V}-n}^+ \otimes 1 \\
+& \sum_{s=0}^{-c_{ij}^{V}-n-1} \binom{-c_{ij}^{V}-n}{s}_{\underline{q}_{ii}}\left( \prod_{r=1}^s (1-\underline{q}_{ii}^{-c_{ij}^{V}-n-r}\underline{q}_{ij}\underline{q}_{ji}) \right) \underline E_i^{s} \otimes \underline E_{j,-c_{ij}^{V}-n-s}^+
\end{align*}
On the other hand, by \eqref{eq:formula delta sub i}, \eqref{eq:Ti(E-)=E+}, \eqref{eq:def Ri} and \eqref{eq:producto coproducto dual},
\begin{align*}
\kappa_{i,j;n}^{-1} \, \mathfrak{R}_i^{\rho_i(V)} & \circ (T_i\otimes T_i)\circ \Delta_i(E_{j,n}^-)= \mathfrak{R}_i^{\rho_i(V)} \left( \underline E_{j,-c_{ij}^{V}-n}^+\otimes 1 + 1 \otimes \underline E_{j,-c_{ij}^{V}-n}^+ \right) \\
&= \underline E_{j,-c_{ij}^{V}-n}^+\otimes 1 + \sum_{s=0}^{-c_{ij}^{V}-n-1} \frac{1}{s_{\underline{q}_{ii}}!} \underline E_i^{s} \otimes (\partial_i^L)^s (\underline E_{j,-c_{ij}^{V}-n}^+),
\end{align*}
so $\underline{\Delta} \circ T_i(E_{j,n}^-) = \mathfrak{R}_i^{\rho_i(V)} \circ (T_i\otimes T_i)\circ \Delta_i(E_{j,n}^-)$.
\end{proof}
\subsection{Coproduct structure of $Z(V)$}
Next we prove that $Z(V)$ is a Hopf subalgebra of $U(V)$. We start by describing the left hand side of the coproduct $\underline{\Delta}$ of $Z^+(V)$. Let $v=s_{i_1}\cdots s_{i_k}\in\Hom(\mathcal{W},V)$ be an element of length $k$, $\beta_j=s_{i_1}\cdots s_{i_{j-1}}(\alpha_{i_j})\in\Delta_+^{V}$, and as above $E_{\beta_j}=T_{i_1}\cdots T_{i_{j-1}}(E_{i_j})$.
\begin{proposition}\label{prop:lado izquierdo Delta Z(chi)}
If $\beta_k\in\mathcal{O}(V)$, then there exist $X(n_1,\dots,n_{k-1})\in U^+(V)$ such that
\begin{align}\label{eq:formula lado izquierdo Delta Z(chi)}
\underline{\Delta}(E_{\beta_k}^{N_{\beta_k}}) &= E_{\beta_k}^{N_{\beta_k}}\otimes 1+1\otimes E_{\beta_k}^{N_{\beta_k}} \\
&\quad + \sum_{n_i\in \mathbb{N}_0} E_{\beta_{k-1}}^{n_{k-1}N_{\beta_{k-1}}} \dots E_{\beta_1}^{n_1N_{\beta_1}} \otimes X(n_1,\dots,n_{k-1}). \notag
\end{align}
\end{proposition}
\begin{proof}
The proof is by induction on $k$. If $k=1$, then $E_{\beta_1}=E_{i_1}$, and $E_{i_1}^{N_{i_1}}$ is primitive.
Assume that \eqref{eq:formula lado izquierdo Delta Z(chi)} holds for every $v'$ of length less than $k$. Let $v=s_{i_1}^{V} v'$, $i=i_1$, $\beta=\beta_k$, $\gamma_j:=s_{i_2}\cdots s_{i_{j-1}}(\alpha_{i_j})\in\Delta_+^{\rho_i(V)}$, $j=2,\cdots,k$, $\gamma=\gamma_k$, so $\beta_j=s_i(\gamma_j)$, $E_{\beta_j}=T_i(E_{\gamma_j})$, $N_\beta=N_\gamma$. By inductive hypothesis,
\begin{align*}
\underline{\Delta}(E_{\gamma}^{N_{\gamma}}) &= E_{\gamma}^{N_{\gamma}}\otimes 1+1\otimes E_{\gamma}^{N_{\gamma}} \\
&\qquad + \sum_{n_i\in \mathbb{N}} E_{\gamma_{k-1}}^{n_{k-1}N_{\gamma_{k-1}}} \dots E_{\gamma_2}^{n_2N_{\gamma_2}}\otimes Y(n_1,\dots,n_{k-1}),
\end{align*}
for some $Y(n_1,\dots,n_{k-1})\in U^+(\rho_i(V))$. As $\partial_i^L(E_{\beta_j}^{N_{\beta_j}})=0$ by Corollary \ref{coro:derivadas PRV}, we have that $E_i\rightharpoonup' E_{\gamma_j}^{N_{\gamma_j}}=0$ by \eqref{eq:Ti con delta L}. By Lemma \ref{lema:caracter sim trivial en Nbeta beta},
\begin{align*}
0 &= \underline{\Delta}(E_i\rightharpoonup' E_{\gamma}^{N_{\gamma}}) = \underline{\Delta}(E_{\gamma}^{N_{\gamma}})(E_i\otimes 1+1\otimes E_i) \\
& \qquad -\chi(N_\gamma \gamma,\alpha_i) (E_i\otimes 1+1\otimes E_i) \underline{\Delta}(E_{\gamma}^{N_{\gamma}}) \\
&= \sum_{n_i\in \mathbb{N}} E_i\rightharpoonup' \left( E_{\gamma_{k-1}}^{n_{k-1}N_{\gamma_{k-1}}} \dots E_{\gamma_2}^{n_2N_{\gamma_2}}\right) \otimes Y(n_2,\dots,n_{k-1}) \\
& \quad + E_{\gamma_{k-1}}^{n_{k-1}N_{\gamma_{k-1}}} \dots E_{\gamma_2}^{n_2N_{\gamma_2}} \otimes E_i\rightharpoonup' Y(n_2,\dots,n_{k-1}) \\
& \quad + \left(1- \prod_{j=2}^{k-1} \chi(N_{\gamma_j}\gamma_j,\alpha_i)^{n_j}\chi(\alpha_i,N_{\gamma_j}\gamma_j)^{n_j} \right)
\chi \left(N_{\gamma}\gamma-\sum_{j=2}^{k-1} N_{\gamma_j}\gamma_j,\alpha_i \right) \\
& \qquad E_{\gamma_{k-1}}^{n_{k-1}N_{\gamma_{k-1}}} \dots E_{\gamma_2}^{n_2N_{\gamma_2}} \otimes Y(n_2,\dots,n_{k-1}) E_i \\
&= \sum_{n_i\in \mathbb{N}} E_{\gamma_{k-1}}^{n_{k-1}N_{\gamma_{k-1}}} \dots E_{\gamma_2}^{n_2N_{\gamma_2}} \otimes E_i\rightharpoonup' Y(n_2,\dots,n_{k-1}),
\end{align*}
so $E_i\rightharpoonup' Y(n_2,\dots,n_{k-1})=0$ for all $n_i\in\mathbb{N}_0$. Moreover $(\underline{\Delta}\otimes \id)\underline{\Delta}(E_{\gamma}^{N_{\gamma}})$ can be written as a sum of terms where the three elements of $U^+(V)^{\otimes 3}$ are products of $E_{\gamma_j}^{n_jN_{\gamma_j}}$ and $Y(n_2,\dots,n_{j-1})$, $1\leq j\leq k$, so they are annihilated by $E_i\rightharpoonup'$. In particular the middle term have terms $E_i^n$ only for $n \in \mathbb{N} N_i$, so the remaining terms in $\iota_{i,V} \mathcal{S}\pi_{i,V}(x_{(2)})$ are of the form $E_i^{nN_i}$, $n\in\mathbb{N}_0$.
As $\Delta_i(E_{\gamma}^{N_{\gamma}})= (\id\otimes m)(\id\otimes\iota_{i,V} \mathcal{S}\pi_{i,V}\otimes\id)(\underline{\Delta}\otimes \id)\underline{\Delta}(E_{\gamma}^{N_{\gamma}})$, there exist $Z(n_2,\dots,n_{k-1})\in U^+(V)$ such that $E_i\rightharpoonup' Z(n_2,\dots,n_{k-1})=0$ and
\begin{align*}
\Delta_i(E_{\gamma}^{N_{\gamma}})& = E_{\gamma}^{N_{\gamma}}\otimes 1+1\otimes E_{\gamma}^{N_{\gamma}} \\
&\qquad + \sum_{n_i\in \mathbb{N}} E_{\gamma_{k-1}}^{n_{k-1}N_{\gamma_{k-1}}} \dots E_{\gamma_2}^{n_2N_{\gamma_2}} \otimes Z(n_2,\dots,n_{k-1}).
\end{align*}
By Theorem \ref{thm:factorization delta Ti},
\begin{align*}
\underline{\Delta}(& E_{\beta}^{N_{\beta}}) = \underline{\Delta}\circ T_i(E_{\gamma}^{N_{\gamma}})=\mathfrak{R}_i^{V} \circ (T_i\otimes T_i)\circ\Delta_i(E_{\gamma}^{N_{\gamma}}) \\
&= \mathfrak{R}_i^{V} \left( E_{\beta}^{N_{\beta}}\otimes 1+1\otimes E_{\beta}^{N_{\beta}} \right) \\
&\qquad + \mathfrak{R}_i^{V} \left( \sum_{n_i\in \mathbb{N}} E_{\beta_{k-1}}^{n_{k-1}N_{\beta_{k-1}}} \dots E_{\beta_2}^{n_2N_{\beta_2}} \otimes T_i(Z(n_2,\dots,n_{k-1})) \right).
\end{align*}
If $n$ is not a multiple of $N_i$, then
\begin{align*}
( T_i & (Z(n_2,\dots,n_{k-1}) ) ) \triangleleft\mathfrak{E}_i^{(n)} \overset{\eqref{eq:producto coproducto dual}}= (n)_{q_{ii}}^{-1} \left( \partial_i^L \circ T_i(Z(n_2,\dots,n_{k-1})) \right)\triangleleft\mathfrak{E}_i^{(n-1)} \\
& \overset{\eqref{eq:Ti con delta L}}= -\chi(\delta,\alpha_i)^{-1}q_{ii}^{-1}(n)_{q_{ii}}^{-1} \left(T_i(E_i\rightharpoonup'Z(n_2,\dots,n_{k-1})) \right)\triangleleft\mathfrak{E}_i^{(n-1)} =0,
\end{align*}
where $\delta=N_\gamma \gamma- \sum_{j=2}^{k-1} n_jN_{\gamma_j}\gamma_j$ is the degree of $Z(n_2,\dots,n_{k-1})$. As also $\delta_i^L(E_{\beta}^{N_{\beta}})=0$, we have that $E_{\beta}^{N_{\beta}}\triangleleft\mathfrak{E}_i^{(n)}=0$. Therefore
\begin{align*}
\underline{\Delta}(& E_{\beta}^{N_{\beta}}) = E_{\beta}^{N_{\beta}} \otimes 1 + \sum_{n\in\mathbb{N}_0} E_i^{n N_i}\otimes E_{\beta}^{N_{\beta}} \triangleleft\mathfrak{E}_i^{(nN_i)} \\
& + \sum_{n_i\in \mathbb{N}} \sum_{n\in\mathbb{N}_0} E_{\beta_{k-1}}^{n_{k-1}N_{\beta_{k-1}}} \dots E_{\beta_2}^{n_2N_{\beta_2}} E_i^{n N_i} \otimes T_i(Z(n_2,\dots,n_{k-1})) \triangleleft\mathfrak{E}_i^{(nN_i)},
\end{align*}
which concludes the inductive step. \end{proof}
\begin{theorem}\label{thm:Z+ es coinv del ker pi}
$Z^+(V)$ is a normal right coideal subalgebra of $U^+(V)$.
Moreover $Z^+(V)=^{co \pi_V|_{U^+(V)}}U^{+}(V).$
\end{theorem}
\begin{proof}
By Proposition \ref{prop:lado izquierdo Delta Z(chi)} $Z^+(V)$ is a right coideal subalgebra, and is normal by Theorem \ref{thm:root vectors almost central, freeness}. Then the right coideal $U^+(V) (Z^+(V))^+$ is a Hopf ideal, and $Z^+(V)=^{co \pi_V|_{U^+(V)}}U^{+}(V)$ by Proposition \ref{pro:masuoka} ${\mathsf {(ii)}\;}$, since $\mathfrak{u}^{\ge 0}(V) = U^+(V)/ U^+(V) (Z^+(V))^+$.
\end{proof}
We consider the following subalgebras of $U^+(V)$:
\begin{align*}
\mathcal{K}^K(V) & =\bigcap_{i=1}^\theta \ker \partial_i^K, & \mathcal{K}^L(V) & =\bigcap_{i=1}^\theta \ker \partial_i^L, & \mathcal{K}(V)&= \mathcal{K}^K(V) \cap\mathcal{K}^L(V).
\end{align*}
\begin{remark}
$\mathcal{K}^L(V)$ is a right coideal subalgebra since it is the intersection of right coideal subalgebras. Similarly $\mathcal{K}^K(V)$ is a left coideal subalgebra.
\end{remark}
\begin{lemma}\label{lema:S entre ker derivaciones}
${\mathsf {(i)}\;}$ For all $x\in U^+(V)$, $i\in\I$, $\partial_i^L(\mathcal{S}(x))= -\mathcal{S}(\partial_i^K(x))$.
\smallbreak
\noindent${\mathsf {(ii)}\;}$ The restriction of $\mathcal{S}$ gives bijective $\mathbb{Z}^{\theta}$-graded antialgebra maps
\begin{align*}
\ker \partial_i^K & \overset{\sim}\longrightarrow \ker \partial_i^L, \quad i\in\I, & \mathcal{K}^K(V) & \longrightarrow \mathcal{K}^L(V).
\end{align*}
\end{lemma}
\begin{proof}
${\mathsf {(i)}\;}$ Let $x\in U^+(V)_n$, $n\in\mathbb{N}_0$. If $n=0$, then both sides are 0. Let $n\ge1$. As $\mathcal{S}$ is an anticoalgebra graded map and $\mathcal{S}(E_i)=-E_i$ for all $i\in\I$,
\begin{align*}
\sum_{i=1}^\theta \partial_i^L(\mathcal{S}(x)) \otimes E_i &= \underline{\Delta}_{1,n-1}^{\cop}\circ\mathcal{S}(x)=(\mathcal{S}\otimes\mathcal{S})\circ\underline{\Delta}_{n-1,1}(x) \\
&= -\sum_{i=1}^\theta \mathcal{S}(\partial_i^K(x)) \otimes E_i.
\end{align*}
\noindent${\mathsf {(ii)}\;}$ From the identity in ${\mathsf {(i)}\;}$, $\mathcal{S}(\ker \partial_i^K)\subseteq \ker \partial_i^L$, $\mathcal{S}(\mathcal{K}^K(V))\subseteq\mathcal{K}^L(V)$. Also $\mathcal{S}$ is a $\mathbb{Z}^{\theta}$-graded, and bijective since $U^+(V)$ is connected.
\end{proof}
We obtain a different characterization of
$Z^+(V)$.
\begin{theorem}\label{thm:Z+ es subalg hopf U+}
The subalgebra $Z^+(V)$ coincides with $\mathcal{K}^L(V)$, $\mathcal{K}^K(V)$. In particular $Z^+(V)$ is a braided Hopf subalgebra of $U^+(V)$.
\end{theorem}
\begin{proof}
By Corollary \ref{coro:derivadas PRV}, $Z^+(V) \subseteq \mathcal{K}^L(V)$. If $x\in \mathcal{K}^L(V)$ is an homogeneous element of degree $n\geq2$, then $\pi_V(x)$ is an homogeneous element of the same degree, annihilated by all the corresponding derivations $\partial_i^L$ of the Nichols algebra $\mathfrak{u}^+(V)$. But the intersection of the kernels of $\partial_i^L$ in $\mathfrak{u}^+(V)$ is $ \mathbf{k} 1$ by \cite[Proposition 2.4]{MS}, so $\pi_V(x)=0$. Then
$ U^+(V) N(\mathcal{K}^L(V))^+ \subseteq \ker\pi_V$. As $\ker\pi_V=U^+(V) (Z^+(V))^+$ by Theorem \ref{thm:Z+ es coinv del ker pi}, both Hopf ideals coincide, so by Proposition \ref{pro:masuoka} $Z^+(V)=N(\mathcal{K}^L(V))$. Therefore $Z^+(V)=\mathcal{K}^L(V)$.
Also $Z^+(V) \subseteq \mathcal{K}^K(V)$ by Corollary \ref{coro:derivadas PRV}, and by Lemma \ref{lema:S entre ker derivaciones}${\mathsf {(ii)}\;}$ the Hilbert series of $\mathcal{K}^K(V)$ and $\mathcal{K}^L(V)$ coincide. As $Z^+(V)=\mathcal{K}^L(V)$, we have that $Z^+(V)=\mathcal{K}^K(V)$. Therefore $Z^+(V)$ is a braided Hopf subalgebra of $U^+(V)$ since it is simultaneously a right and a left coideal subalgebra.
\end{proof}
\begin{corollary}
The subalgebras $Z^\pm(V)$, $Z^0(V)$, $Z(V)$ do not depend on the choice of the element of maximal length.
\end{corollary}
\begin{proof}
By Theorem \ref{thm:Z+ es subalg hopf U+}, $Z^+(V)=\mathcal{K}(V)$, and $\mathcal{K}(V)$ does not depend on that element of maximal length; the proof for $Z^-(V)$ is analogous. As $Z^0(V)$ is the group algebra of the subgroup generated by $K_\beta^{N_\beta}$, $L_\beta^{N_\beta}$, $\beta\in\mathcal{O}(V)$, the proof for $Z(V)$ follows since it is generated by $Z^\pm(V)$, $Z^0(V)$.
\end{proof}
We now deduce the corresponding statement for $Z(V)$ and $U(V)$.
\begin{theorem}\label{thm:Z es subalg hopf U}
$Z(V)$ is a Hopf subalgebra of $U(V)$.
\end{theorem}
\begin{proof}
Let $\beta\in\mathcal{O}(V)$. The $U^{+0}(V)$-coaction is $\lambda(E_\beta^{N_\gamma})=K_\gamma^{N_\gamma}\otimes E_\gamma^{N_\gamma}$, $\gamma\in\mathcal{O}(V)$, and $\underline{\Delta}(Z^+(V))\subset Z^+(V)\otimes Z^+(V)$ by Theorem \ref{thm:Z+ es subalg hopf U+}, so $\Delta(Z^+(V))\subset Z^+(V)\Gamma(V)\otimes Z^+(V)$. Analogously, $\Delta(Z^-(V))\subset Z^-(V)\otimes Z^-(V)\Gamma(V)$.
\end{proof}
\begin{remark}
Assume that $\chi$ is such that $\chi(N_\beta\beta,\alpha_j)$=1 for all $\beta\in\mathcal{O}(V)$, $1\le j\le\theta$. Then $Z(V)$ is a central Hopf subalgebra and $\Gamma(V)$ acts trivially on $U^\pm(V)$, $\mathfrak{u}^\pm(V)$ so the quotient
\begin{align*}
\widehat{\mathfrak{u}}(V)=\mathfrak{u}(V)/<K^{N_\beta\beta}-1,L^{N_\beta\beta}-1:\beta\in\mathcal{O}(V)>
\end{align*}
is well defined. The triangular decomposition on $\mathfrak{u}(V)$ induces a triangular decomposition on $\widehat{\mathfrak{u}}(V)$, where $\widehat{\mathfrak{u}}^\pm(V)$ is canonically identified with $\mathfrak{u}^\pm(V)$, and the zero part is $\widehat{\mathfrak{u}}^0(V)\simeq \mathbf{k}(\mathbb{Z}^{2\theta}/\Gamma(V))$. The algebra $\widehat{\mathfrak{u}}(V)$ can be seen as a quantum double of the bosonizations of $\mathfrak{u}^\pm(V)$ with $ \mathbf{k}(\mathbb{Z}^{\theta}/<K^{N_\beta\beta}-1:\beta\in\mathcal{O}(V)>)$. We have an extension of Hopf algebras \cite{AD}:
$$ \mathbf{k} \rightarrow Z(V) \hookrightarrow U(V) \twoheadrightarrow \widehat{\mathfrak{u}}(V) \rightarrow \mathbf{k}, $$
which generalizes the case of quantum groups $U_q(\mathfrak g)$ at roots of unity and the corresponding small quantum groups $\mathfrak{u}_q(\mathfrak g)$ obtained as a quotient.
\end{remark}
\begin{remark}
The graded dual $\mathcal{L}(V)=U^+(V)^*$ is a braided Hopf algebra containing a copy of $\mathfrak{u}^+(V)^* \simeq \mathfrak{u}^+(V)$; moreover is a finite module over this subalgebra. $\mathcal{L}(V)$ was called the \emph{Lusztig algebra} \cite{AAGTV} and resembles the $q$-divided power algebra \cite{L-libro}.
\end{remark}
\section{Some examples}\label{section:examples}
\subsection{Braidings of super type $A$}
We apply previous results to braidings of super type $A$, see \cite{AAY}. Fix $N\in\mathbb{N}$, $q\in \mathbf{k}$ a root of unity of order $N\geq 3$. There exists a braiding $C(\theta,q;i_1,\ldots ,i_j)$ of super type $A_\theta$ for each ordered subset $1\leq i_1<\dots<i_k\leq\theta$. Its matrix $(q_{ij})_{1 \leq i,j \leq \theta}$ satisfies
\begin{itemize}
\smallbreak\item $q_{ij}q_{ji}=1$ if $1 \le i, j<\theta$, $|i-j|>1$,
\smallbreak\item if $i=i_\ell$ for some $1\leq\ell\le k$, then $q_{ii}=-1$, $q_{i-1,i}q_{i,i-1}q_{i+1,i}q_{i,i+1}=1$,
\smallbreak\item if $i\neq i_\ell$ for all $1\leq\ell\le k$, then $q_{ii}q_{i-1,i}q_{i,i-1}=q_{ii}q_{i+1,i}q_{i,i+1}=1$,
\end{itemize}
where $q=q_{11}^2 q_{12}q_{21}$. The notation is close to the simple chains in \cite{H-classif RS}.
The positive roots corresponding to this braiding are
\begin{align*}
\Delta_+^{V} &= \{ \alpha_{j,k}:1\le j\le k\le\theta\}, & \mbox{where } \alpha_{j,k} & =\alpha_j+\alpha_{j+1} +\dots + \alpha_k.
\end{align*}
Set $t_k^{V}=s_1^{V} s_2\cdots s_k$, $w=t_\theta^{V} t_{\theta-1}\dots t_1$, where we use for the $t$'s the same convention for concatenation as for the $s$'s. The expression of $w$ in $s_i$'s is reduced and this is the element of maximal length by Lemma \ref{Lemma:longitudHY} since
\begin{align}
t_\theta &t_{\theta-1}\dots t_{\theta-j+2} s_1\dots s_{k-j-1}(\alpha_{k-j})= \alpha_{j,k} & &1\le j\le k\le\theta.
\end{align}
Let $| \, |:\mathbb{Z}^\theta\to\mathbb{Z}_2$ be the group map such that $|\alpha_i|=1$ if $i\in\{i_1,\dots,i_k\}$ and $|\alpha_i|=0$ otherwise, which defines the parity of the roots. $\mathcal{O}(V)$ is the set of even roots, $\chi(\beta,\beta)=q^{\pm1}$ if $\beta\in\mathcal{O}(V)$, and $\chi(\beta,\beta)=-1$ otherwise.
The algebra $U^+(V)$ is presented by generators $E_i$, $1\le i\le\theta$, and relations
\begin{align}
&E_iE_j=q_{ij} E_jE_i, & &j-i \geq 2, \label{eq:relaciones U tipo A 1} \\
&[(\ad_c E_{i-1})(\ad_c E_i) E_{i+1}, E_i]_c, \quad E_i^2, & i &\in\{i_1,\dots,i_k\}, \label{eq:relaciones U tipo A 2} \\
&(\ad_c E_i)^2 E_{i\pm1}, & i & \notin\{i_1,\dots,i_k\}. \label{eq:relaciones U tipo A 3}
\end{align}
Let $E_{j,k} = (\ad_c E_j)\cdots (\ad_c E_{k-1})E_k= T_j\cdots T_{k-1}(E_k)$, $1\le j\le k\le\theta$. They are the generators of the PBW basis corresponding to the previous expression of $w$, ordered lexicographically:
$$ \alpha_{1,1}<\alpha_{1,2} <\cdots <\alpha_{1,\theta} <\alpha_{2,2} <\alpha_{2,3} <\cdots <\alpha_{\theta,\theta}. $$
By Remark \ref{rem:PRV=0 si alpha no cartan} $E_{j,k}^2=0$ for each odd root $\alpha_{j,k}$.
The computation of $\underline{\Delta}(E_{j,k})$ follows as for braidings of Cartan type with matrix $A_\theta$, see \cite[Lemma 6.5]{AS Pointed HA}.
\begin{lemma}\label{lema:coproducto Ejk, tipo A}
Let $1\le j\le k\le\theta$. Then
\begin{align}\label{eq:coproducto Ejk, tipo A}
\underline{\Delta}(E_{j,k})= E_{j,k}\otimes 1+1\otimes E_{j,k}+ \sum_{j\le \ell<k}(1-\widetilde{q_{l,l+1}}) E_{j,\ell} \otimes E_{\ell+1,k}.
\end{align}
\end{lemma}
\begin{proof}
The proof is by induction on $k-j$, the case $k=j$ is trivial. Assume that \eqref{eq:coproducto Ejk, tipo A}
holds for $j+1$, $k$.
Notice that
\begin{align*}
\chi(\alpha_j,\alpha_{j+1,k})\chi(\alpha_{j+1,k},\alpha_j)&=\widetilde{q_{j,j+1}},&
\chi(\alpha_j,\alpha_{\ell+1,k})\chi(\alpha_{\ell+1,k},\alpha_j)&=1 \mbox{ if }\ell>j.
\end{align*}
Then we compute:
\begin{align*}
\underline{\Delta}(E_{j,k})& = \underline{\Delta}(E_j)\underline{\Delta}(E_{j+1,k})-\chi(\alpha_j,\alpha_{j+1,k}) \underline{\Delta}(E_{j+1,k})\underline{\Delta}(E_j) \\
&= E_{j,k}\otimes 1+1\otimes E_{j,k}+ (1-\widetilde{q_{j,j+1}}) E_j\otimes E_{j+1,\ell} \\
& \qquad + \sum_{j+1\le \ell<k} (1-\widetilde{q_{l,l+1}}) \, E_{j,\ell} \otimes E_{\ell+1,k}\\
& \qquad + \sum_{j+1\le \ell<k} (1-\widetilde{q_{l,l+1}})\chi(\alpha_j,\alpha_{j+1,k}) E_{j+1,\ell} \otimes [E_j, E_{\ell+1,k}]_c,
\end{align*}
From \eqref{eq:relaciones U tipo A 1}, $[E_j, E_{\ell+1,k}]_c=0$ if $\ell>j$, which completes the proof.
\end{proof}
We compute now $\underline{\Delta}(E_{j,k}^N)$ for each even root $\alpha_{j,k}$. Compare this with the case of Cartan braidings of type $A$ \cite[Lemma 6.9]{AS Pointed HA}.
\begin{proposition}
Let $1\le j\le k\le\theta$ be such that $\alpha_{j,k}\in\mathcal{O}(V)$. Then
\begin{align}
\underline{\Delta}&(E_{j,k}^N)= E_{j,k}^N\otimes 1+1\otimes E_{j,k}^N \notag\\
& + \sum_{\ell: \, j\le \ell<k, \, \alpha_{j,\ell}\in\mathcal{O}(V)} (1-\widetilde{q_{\ell,\ell+1}})^N \chi(\alpha_{j,\ell},\alpha_{\ell+1,k})^{\frac{N(N-1)}{2}} E_{j,\ell}^N \otimes E_{\ell+1,k}^N.
\end{align}
\end{proposition}
\begin{proof}
By induction on $k-j$. The case $k=j$ is trivial. Assume that
\begin{align*}
\underline{\Delta}&(E_{j+1,k}^N)= E_{j+1,k}^N\otimes 1+1\otimes E_{j+1,k}^N \\
& + \sum_{\underset{\alpha_{j+1,\ell}\in\mathcal{O}(\rho_j(V))} {\ell: \, j+1\le \ell<k,} } (1-\widetilde{\underline{q}_{\ell,\ell+1}})^N s_j^*\chi(\alpha_{j+1,\ell},\alpha_{\ell+1,k})^{\frac{N(N-1)}{2}} E_{j+1,\ell}^N \otimes E_{\ell+1,k}^N.
\end{align*}
Then $ \Delta_j(E_{j+1,k}^N)=\underline{\Delta}(E_{j+1,k}^N)$ since $(\underline{\Delta}\otimes\id)\underline{\Delta}(E_{j+1,k}^N)$ does not have terms $E_j^n$, $n\in\mathbb{N}$.
If $\ell>j$, then $\widetilde{\underline{q}_{\ell,\ell+1}}= \widetilde{q_{\ell,\ell+1}}$, and
\begin{itemize}
\item $\alpha_{j+1,\ell}\in\mathcal{O}(\rho_j(V))$ if and only if $s_j(\alpha_{j+1,\ell})=\alpha_{j,\ell}\in\mathcal{O}(V)$,
\item $s_j^*\chi(\alpha_{j+1,\ell},\alpha_{\ell+1,k})=\chi(s_j(\alpha_{j+1,\ell}),s_j(\alpha_{\ell+1,k}))=\chi(\alpha_{j,\ell},\alpha_{\ell+1,k})$.
\end{itemize}
We apply now Theorem \ref{thm:factorization delta Ti}:
\begin{align*}
\underline{\Delta}(E_{j,k}^N & )= \underline{\Delta}\circ T_j (E_{j+1,k}^N) = \mathfrak{R}_i^{\rho_j(V)}\circ(T_j\otimes T_j)\circ \Delta_j(E_{j+1,k}^N) \\
&=
\mathfrak{R}_i^{\rho_j(V)} \Big( E_{j,k}^N\otimes 1+1\otimes E_{j,k}^N\Big) \\
& + \sum_{\underset{\alpha_{j,\ell}\in\mathcal{O}(V)} {\ell: \, j+1\le \ell<k,} } (1-\widetilde{q_{\ell,\ell+1}})^N \chi(\alpha_{j,\ell},\alpha_{\ell+1,k})^{\frac{N(N-1)}{2}} \mathfrak{R}_i^{\rho_j(V)} \Big(E_{j,\ell}^N \otimes E_{\ell+1,k}^N \Big).
\end{align*}
$\mathfrak{R}_i^{\rho_j(V)}\Big(E_{j,\ell}^N \otimes E_{\ell+1,k}^N \Big)=\Big(E_{j,\ell}^N \otimes E_{\ell+1,k}^N \Big)$ since $E_{\ell+1,k}^N \triangleleft\mathfrak{E}_j^{(n)} =0$ for all $n\geq 1$, $\ell>j$. Also $\mathfrak{R}_i^{\rho_j(V)} \left( E_{j,k}^N\otimes 1 \right)=E_{j,k}^N\otimes 1$. By Theorem \ref{thm:Z+ es subalg hopf U+} it suffices to compute $E_{j,k}^N \triangleleft\mathfrak{E}_j^{(nN)}$. As $\underline{\Delta}$ is $\mathbb{N}_0$-graded, $E_{j,k}^N \triangleleft\mathfrak{E}_j^{(nN)}=0$ if $n>1$, so we compute $X_N$, where $E_j^N\otimes X_n$ is a summand in the expression of $\underline{\Delta}(E_{j,k}^N)=\underline{\Delta}(E_{j,k})^N$. By Lemma \ref{lema:coproducto Ejk, tipo A},
\begin{align*}
E_j^N\otimes X_n & = (1-\widetilde{q_{j,j+1}})^N (E_j \otimes E_{j+1,k})^N \\
& = (1-\widetilde{q_{j,j+1}})^N \chi(\alpha_j,\alpha_{j+1,k})^{\frac{N(N-1)}{2}} \, E_j^N \otimes E_{j+1,k}^N,
\end{align*}
so we have
$$ \mathfrak{R}_i^{\rho_j(V)} \left( 1\otimes E_{j,k}^N \right)= 1\otimes E_{j,k}^N + (1-\widetilde{q_{j,j+1}})^N \chi(\alpha_j,\alpha_{j+1,k})^{\frac{N(N-1)}{2}} \, E_j^N \otimes E_{j+1,k}^N, $$
which completes the proof. \end{proof}
\subsection{Braidings of type $\mathfrak{br}(2;5)$}
Now fix $\theta=2$, $\zeta$ a root of unity of order 5, $(q_{ij})_{i,j\in\I}$, $(r_{ij})_{i,j\in\I}$ two matrices such that
\begin{align*}
q_{11}&=\zeta, & \widetilde{q_{12}}&=\zeta^2, & q_{22}&=r_{22}=-1, & r_{11}&=-\zeta^3, & \widetilde{r_{12}}&=\zeta^3.
\end{align*}
Let $\chi,\psi:\mathbb{Z}^2\times\mathbb{Z}^2\to \mathbf{k}^\times$ be the bicharacters defined by $(q_{ij})_{i,j\in\I}$, $(r_{ij})_{i,j\in\I}$, respectively, and $V$, $W$ the corresponding braided vector spaces.
They correspond to \cite[Table 1, row 13]{H-classif RS} and are related with the Lie superalgebra $\mathfrak{br}(2;5)$ over a field of characteristic 5 \cite{AA-WGCLSandNA}. Their generalized Cartan matrices are
$ C^{V}= \left( \begin{array}{cc} 2 & -3 \\ -1 & 2 \end{array}\right)$, $ C^W= \left( \begin{array}{cc} 2 & -4 \\ -1 & 2 \end{array}\right)$, and $\rho_1(V)=V$, $\rho_1(W)=W$, $\rho_2(V)=W$. The elements of maximal length are $(\sigma_1^{V}\sigma_2)^4$, $(\sigma_1^W\sigma_2)^4$, and the positive roots $\Delta_+^{V}$, $\Delta_+^W$ are, respectively,
\begin{align*}
\big\{ \alpha_1,3\alpha_1+\alpha_2,2\alpha_1+\alpha_2,5\alpha_1+3\alpha_2,3\alpha_1+2\alpha_2,4\alpha_1+3\alpha_2,\alpha_1+\alpha_2,
\alpha_2 \big\}, \\
\big\{\alpha_1,4\alpha_1+\alpha_2,3\alpha_1+\alpha_2,5\alpha_1+2\alpha_2,2\alpha_1+\alpha_2,3\alpha_1+2\alpha_2,\alpha_1+\alpha_2,
\alpha_2 \big\}.
\end{align*}
The sets of Cartan roots $\mathcal{O}(V)$, $\mathcal{O}(W)$ are, respectively,
\begin{align*}
&\left\{\alpha_1,2\alpha_1+\alpha_2,3\alpha_1+2\alpha_2,\alpha_1+\alpha_2\right\}, &
&\left\{\alpha_1,3\alpha_1+\alpha_2,2\alpha_1+\alpha_2,\alpha_1+\alpha_2 \right\}.
\end{align*}
The algebra $U^+(V)$ is presented by generators $E_1,E_2$ and relations
\begin{align*}
&E_2^2, & &(\ad_cE_1)^4E_2, & &[E_{1112},E_{112}]_c, & &\left[E_{4\alpha_1+3\alpha_2},E_{12}\right]_c.
\end{align*}
The algebra $U^+(W)$ is presented by generators $\underline E_1,\underline E_2$ and relations
\begin{align*}
&\underline E_2^2, & &(\ad_c\underline E_1)^5\underline E_2, & &\left[\underline E_1,\underline E_{3\alpha_1+2\alpha_2}\right]_c+r_{12}\underline E_{2\alpha_1+\alpha_2}^2, & &\left[\underline E_{3\alpha_1+2\alpha_2},\underline E_{12}\right]_c.
\end{align*}
Here $E_{i_1\cdots i_kj}=(\ad_c E_{i_1})\dots(\ad_c E_{i_{k}})E_j$, and analogous notation with $\underline E$.
\begin{lemma}\label{lema:root vectors br(2,5)}
The root vectors of $\mathcal{O}(V)$, $\mathcal{O}(W)$ are, respectively,
\begin{align*}
E_{\alpha_1}&=E_1, & E_{3\alpha_1+2\alpha_2}&= q_{12}^3\zeta^3(1-\zeta^3)^5(1+\zeta) [E_{112},E_{12}]_c, \\
E_{2\alpha_1+\alpha_2}&=q_{21}(1-\zeta^3)E_{112}, & E_{\alpha_1+\alpha_2}&=q_{21}^2\zeta(1-\zeta^3)^7(1+\zeta)^3 E_{12}, \\
\underline E_{\alpha_1}&=\underline E_1, & \underline E_{2\alpha_1+\alpha_2}&= -r_{12}^2\zeta^2(1+\zeta)^2(1-\zeta^3)^4 \underline E_{112}, \\
\underline E_{3\alpha_1+\alpha_2}&=r_{21}(1-\zeta^2)\underline E_{1112}, & \underline E_{\alpha_1+\alpha_2}&=r_{12}^2\zeta(1+\zeta)^3(1-\zeta^3)^{10} \underline E_{12}.
\end{align*}
\end{lemma}
\begin{proof}
First we compute
\begin{align*}
E_{2\alpha_1+\alpha_2}&=T_1^{V} T_2(\underline E_1)=T_1^{V}\left( E_{21} \right) = -q_{21}\zeta \partial_1^L( E_{1112})=q_{21}(1-\zeta^3)E_{112}.
\end{align*}
A similar computation gives the expression of $\underline E_{3\alpha_1+\alpha_2}$. Using this,
\begin{align*}
E_{3\alpha_1+2\alpha_2} &=T_1^{V} T_2 (\underline E_{3\alpha_1+\alpha_2}) =(1-\zeta^2)^2 T_1^{V}([[E_{21},[E_{21},E_1]_c]_c) \\
&= [-L_1\cdot \partial_1^L(E_{112}), L_1\cdot \partial_1^L(L_1\cdot \partial_1^L(E_{112}))]_c\\
&= q_{12}^3\zeta^3(1-\zeta^3)^5(1+\zeta) [E_{112},E_{12}]_c.
\end{align*}
The proof for the remaining root vectors is analogous by iterating compositions of $T_1^{V} T_2$ and $T_1^W
T_2$.
\end{proof}
Now we compute the coproduct of powers of root vectors.
\begin{lemma}\label{lemma:coproducto br(2,5)}
$E_{\alpha_1}^5, E_{\alpha_1+\alpha_2}^{10}$, $\underline E_{\alpha_1}^{10}, \underline E_{\alpha_1+\alpha_2}^5$ are primitive, and
\begin{align*}
\underline{\Delta}(E_{2\alpha_1+\alpha_2}^{10})&= E_{2\alpha_1+\alpha_2}^{10}\otimes 1+1\otimes E_{2\alpha_1+\alpha_2}^{10} + \frac{q_{21}^{30}E_{\alpha_1}^{10}\otimes E_{\alpha_1+\alpha_2}^{10}}{(1-\zeta^3)^{55}(1+\zeta)^{25}} \\
& \qquad \qquad + \frac{q_{21}^{30}}{(1-\zeta^3)^{10}} E_{\alpha_1}^5 \otimes E_{3\alpha_1+2\alpha_2}^5, \\
\underline{\Delta}(E_{3\alpha_1+2\alpha_2}^5)&= E_{3\alpha_1+2\alpha_2}^5 \otimes 1+1\otimes E_{3\alpha_1+2\alpha_2}^5+ \frac{q_{12}^{15}E_{\alpha_1}^5 \otimes E_{\alpha_1+\alpha_2}^{10}}{(1-\zeta^3)^{40}(1+\zeta)^{15}} ,\\
\underline{\Delta}(\underline E_{3\alpha_1+\alpha_2}^5)&= \underline E_{3\alpha_1+\alpha_2}^5 \otimes 1+1\otimes \underline E_{3\alpha_1+\alpha_2}^5 + \frac{r_{21}^{35}\zeta^2 \underline E_{\alpha_1}^{10} \otimes \underline E_{\alpha_1+\alpha_2}^5}{(1-\zeta^3)^{40}(1+\zeta)^5} , \\
\underline{\Delta}(\underline E_{2\alpha_1+\alpha_2}^{10})&= \underline E_{2\alpha_1+\alpha_2}^{10} \otimes 1+1\otimes \underline E_{2\alpha_1+\alpha_2}^{10} - \frac{r_{21}^{45} \underline E_{\alpha_1}^{10} \otimes \underline E_{\alpha_1+\alpha_2}^{10}}{(1-\zeta^3)^{50}} \\
& \qquad \qquad - \frac{r_{12}^5(1+\zeta)^{10}}{(1-\zeta^3)^{10}} \underline E_{3\alpha_1+\alpha_2}^5 \otimes \underline E_{\alpha_1+\alpha_2}^5.
\end{align*}
\end{lemma}
\begin{proof}
We know that $E_{\alpha_1}^5$, $\underline E_{\alpha_1}^{10}$ are primitive. Now $E_{\alpha_1+\alpha_2}^{10}$, $\underline E_{\alpha_1+\alpha_2}^5$ are also primitive since $\underline{\Delta}$ is $\mathbb{N}_0^\theta$-graded and $Z^+(V)$ is a Hopf subalgebra by Theorem \ref{thm:Z+ es subalg hopf U+}. Also $\underline{\Delta}(E_{3\alpha_1+2\alpha_2}^5)$ is the sum of the two terms $\underline E_{3\alpha_1+\alpha_2}^5 \otimes 1$, $1\otimes \underline E_{3\alpha_1+\alpha_2}^5$ plus $c \,E_{\alpha_1}^5 \otimes E_{\alpha_1+\alpha_2}^{10}$ for some $c\in \mathbf{k}$ since the first term of the coproduct should be a power of previous roots by Proposition \ref{prop:lado izquierdo Delta Z(chi)} and then we use again Theorem \ref{thm:Z+ es subalg hopf U+}. As
$$ \underline{\Delta}([E_{112},E_{12}]_c^5)=(1\otimes [E_{112},E_{12}]_c +E_1\otimes \partial_1^L([E_{112},E_{12}]_c)+... )^5,$$
and $\partial_1^L([E_{112},E_{12}]_c)=-\zeta^3(1-\zeta^3)(1+\zeta)^2 E_{12}$, such $c$ is computed from $(E_1 \otimes E_{12}^2)^5=q_{21}^{10} E_1^5\otimes E_{12}^{20}$ using Lemma \ref{lema:root vectors br(2,5)}.
Similarly $\underline{\Delta}(E_{2\alpha_1+\alpha_2}^{10})$ can have two extra summands $a E_{\alpha_1}^{10}\otimes E_{\alpha_1+\alpha_2}^{10}$ and $b E_{\alpha_1}^5 \otimes E_{3\alpha_1+2\alpha_2}^5$. As $\partial_i^L(E_{112})=(1+\zeta)(1-\zeta^3)E_{12}$, the first appears from
$$ (E_1\otimes E_{12})^{10} = q_{21}^{45} E_1^{10}\otimes E_{12} ,$$
while the second appears from the product
$$ ((1\otimes E_{112})(E_1\otimes E_{12}))^{5} = q_{21}^{15} E_1^5 \otimes (E_{112}E_{12})^5 ,$$
The calculus of coproducts of the other two expressions is similar.
\end{proof}
\begin{remark}\label{rem:accion br(2,5)}
The coproduct of the power root vectors in Lemma \ref{lemma:coproducto br(2,5)} seems close to the algebra of functions of $B_2$, and the Lie algebra $\mathfrak g_0$ corresponding to the Lie superalgebra $\mathfrak{br}(2;5)$ in characteristic 5 is of type $B_2$. The PBW bases of $U(V)$ and $\mathfrak{u}(V)$ resemble the PBW bases of the enveloping and the restricted enveloping algebras of the Lie superalgebra of $\mathfrak{br}(2;5)$ respectively, since the generators have the same degree and the corresponding height.
\end{remark}
|
\section{Introduction}\label{sect:intro}
Stellar spectroscopy of stars with spectral types F, G and K is a mature field of astrophysics. At least for simple atoms like lithium and sodium, the matter-light interaction taking place in stellar atmospheres can nowadays be modelled from first principles, with due consideration of hydrodynamics and departures from local thermodynamic equilibrium (LTE). Using long-lived stars as data carriers of different phases of Galactic evolution, much has been learned about the matter cycle and the general build-up of the Galactic inventory of chemical elements with time. However, our inferences are not free from uncertainties and limitations. One such inquietude is the theoretical expect that the surface abundances of late-type stars with shallow outer convection zones are a direct function of time, due to element-separating effects collectively referred to as atomic diffusion \citep{Michaud1984}. Subject to the prevailing forces (gravity, radiation pressure), chemical elements will be pushed into/out of the atmosphere, leaving the visible surface somewhat enriched/depleted, to various degrees for different elements. Given that convection can efficiently counteract atomic diffusion, the effects are expected to lead to sizable surface effects only in hotter stars, i.e. in the context of this paper the F-type turnoff-point stars. During prior and subsequent evolution, the deep outer convection zone retains/restores the surface abundances (with the notable exception of lithium which is mostly destroyed during the course of the first dredge-up on the red-giant branch).
In this series of papers, we investigate to which extent careful spectroscopic studies can trace small yet systematic abundance differences between groups of stars representing different phases in the evolution of globular-cluster stars. The basic idea is simple: with sufficiently accurate spectroscopic modelling, inferred abundance differences can be compared to the predictions of stellar-structure models including the effects of atomic diffusion.
Earlier papers in the series have traced such effects in NGC\,6397 at $\left[\mbox{Fe/H}\right] \footnote{We adopt here the usual spectroscopic notations that $\left[\mbox{X/Y}\right] \equiv \log\,(N_{\rm X}/N_{\rm Y})_{*} - \log\,(N_{\rm X}/N_{\rm Y})_{\odot}$, and that $\log\,\varepsilon({\rm X})\equiv \log\,(N_{\rm X}/N_{\rm H})+12$ for elements X and Y. We assume also that metallicity is equivalent to the stellar [Fe/H] value.} = -2.1$ \citep{Korn2007, Lind2008, Nordlander2012} and NGC\,6752 at $\left[\mbox{Fe/H}\right] = -1.6$ \citep[][hereafter Paper~I]{Gruyters2013}. It was shown that the turnoff-point (TOP) stars do show systematically lower abundances (by 0.05-0.2\,dex) than the red-giant branch (RGB) stars, with intermediate groups falling in between. But the predictions of atomic diffusion are not supported by these observations straight off. To achieve agreement between theory and observation, atomic diffusion needs to be counteracted by what is referred to as additional mixing, i.e.\ non-canonical mixing not provided by convection in the mixing-length formulation. Some mixing beyond the formal extent of the convection zone in the turnoff-point phase of evolution is also needed to explain the properties of the Spite plateau of lithium \citep{Spite1982} which at a given metallicity is found to be mono-valued and not a function of effective temperature \citep{Richard2005}. Such mixing can be prescribed in stellar-structure models by an extra term in the diffusion equation, adding a simple analytical function that accomplishes mixing down to layers of a specific temperature. The free parameter involved in this modelling, the efficiency relative to atomic diffusion, has to, at present, be determined empirically from the observations, i.e.\ from the amplitude of the elemental abundance trends.
For stars in NGC\,6397, the efficiency of additional mixing is well-constrained and falls within the range that prevents lithium destruction which would be incompatible with the properties of the Spite plateau (see above). The agreement between theory and observation is very good and succeeds in describing abundance structures seen for the first time in lithium and calcium \citep{Korn2006, Lind2009}. There is, however, no complete consensus on the effective-temperature scale to be applied, with consequences for the abundance trends of e.g.\ lithium between main-sequence (MS) and sub-giant-branch (SGB) stars \citep{Gonzalez2009b}.
For stars in NGC\,6752, the results are less compelling, mainly because the abundance trends are weaker in this cluster. Nonetheless, with somewhat more efficient additional mixing, the trends can be reproduced and the diffusion-corrected lithium abundance raised to levels compatible with WMAP-calibrated predictions of standard Big-Bang Nucleosynthesis \citep{Gruyters2013}.\\
In this paper, we revisit NGC\,6752 with additional observations obtained with the GIRAFFE spectrograph at the VLT. With improved stellar statistics, we can address questions and potential biases associated with the various stellar populations that make up NGC\,6752. The paper is organized in the following manner: Section 2 describes the observations, both photometric and spectroscopic, and the data reduction providing the material for the analysis presented in Section 3. Section 4 describes the results, while Section 5 discusses their scientific significance. In Section 6, the results are summarized and some conclusions are drawn.
\section{Observations and Data Reduction}\label{sect:obs}
\subsection{Photometry}
Photometric data was obtained using the DFOSC instrument on the 1.5\,m telescope on La Silla, Chile, in 1997 and consist of $uvby$ Str\"omgren photometry. The data is the same as used in Paper~I, and details on the data reduction and photometric calibration procedures can be found in \citet{Grundahl1998,Grundahl1999}.
The sample used for spectroscopic analysis consists of 194 stars located between the cluster turnoff point (TOP) at $V=17.0$ and the base of the red giant branch (bRGB) up to $V=15.3$. A colour-magnitude diagram (CMD) of the total sample is given in Fig.~\ref{Fig:CMD}. We construct colour-magnitude fiducial sequences by averaging the colours of cluster members along the sequence. By interpolating colours at constant $V$-magnitude onto this sequence, we effectively remove the observational scatter, with residual errors dominated by uncertainties in the shape of the fiducial sequence. The best-fitting fiducial is found by trial and error. We find the most reliable method of constructing colour-magnitude fiducials to be quadratic interpolation between points placed manually such that the resulting residuals along the sequence were minimized. Finally, we apply a reddening correction of $E(B-V) = 0.04$ \citep[latest web update]{Harris1996} with the relation coefficients given by \citet{Ramirez2005}, and $E(v-y)=1.7\times E(b-y)$.
\begin{figure}
\begin{center}
\includegraphics{CMD_vy.eps}
\caption{Observed colour-magnitude diagram of NGC 6752. The spectroscopic targets are marked by black squares for the GIRAFFE sample. For comparison we have included the UVES targets from \citet{Gruyters2013} as red diamonds.}\label{Fig:CMD}
\end{center}
\end{figure}
\subsection{Spectroscopy}
The spectroscopic observations were obtained in Service Mode during ESO period 79 (Program ID: 079.D-0645(A)) and period 81 (Program ID: 081.D-0253(A)) using the multi-object spectrograph FLAMES/GIRAFFE mounted on the ESO VLT-UT2 Kueyen \citep{Pasquini2003}. Running in MEDUSA mode, FLAMES-GIRAFFE can simultaneously observe 132 objects. On average we dedicated 15 of the 132 fibres to a simultaneous monitoring of the sky background. As our faintest targets reach $V \approx 17$, we typically coadded multiple long exposures (up to $4215\,\text s$) in order to reach spectra of sufficient quality with signal-to-noise ratios per pixel $(S/N)$ above 30. Including overhead, we spent a total of $70$\,h under good seeing conditions ($\sim 0.7''$) during grey-to-bright lunar time periods using six high-resolution ``B''-settings (HR9B, HR10B, HR11B, HR13B, HR14B, and HR15B).
\begin{table*}
\caption{Observational data of the NGC\,6752 stars observed with FLAMES-GIRAFFE.}
\label{Tab:log}
\centering
\begin{tabular}{lcccccc}
\hline\hline
Setting & Wavelength & Resolution & $t_{\mbox{\scriptsize exp}}$ & $N_\text{stars}$ & ESO period & S/N \\
\# & range (\AA) & $\lambda/\Delta\lambda$ & (hr) & & & Min-Median-Max \\
\hline
HR9 & 5143-5356 & 25900 & 12.9 & 98 & P81 & 38-65-134 \\
HR10 & 5339-5619 & 19800 & 3.5 & 99 & P81 & 26-53-84 \\
HR11 & 5597-5840 & 24200 & 8.4 & 138 & P79 \& P81 & 18-61-104 \\
HR15 & 6607-6965 & 19300 & 4.5 & 126 & P79 & 30-87-142 \\
\hline
\end{tabular}
\tablefoot{Only data analysed in this work are listed.}
\end{table*}
In this work we present the data of four of the six settings, covering a combined spectral range of 1100\,\AA. Stars in different evolutionary stages are distributed randomly across the field-of-view. We find no gradient in background light over the field and subtract a single, median averaged sky-spectrum from the stellar spectra obtained during each observation. The setup is summarised in Table~\ref{Tab:log} for the subset of spectra analysed in this work.
As UVES targets were prioritized during observations, fibre positioning limitations resulted in only 37 stars out of the full sample of 194 being observed in all six settings. The largest overlapping sample consists of 105 stars observed in HR11B covering e.g. the \ion{Mg} i 5711\,\AA\ line and the \ion{Na} i 5682-5688\,\AA\ doublet, and HR15B covering the \ion{Li} i 6707\,\AA\ line and the \ion{Al} i 6696-6698\,\AA\ doublet. Additionally, 99 stars were observed in both HR10B and HR11B, the former covering the \ion{Mg} i 5528\,\AA\ line.
We executed basic data reduction on the spectroscopic data using the GIRAFFE pipeline maintained by ESO.
Individual exposures, ranging between 1 and 11 per star, were coadded after sky-subtraction, normalization, radial-velocity correction and cosmic-ray rejection. As we have only a single exposure for some stars, we opted for a threshold-rejection method of affected pixels and their neighbours, and visually inspected all spectra before and after coaddition.
The minimum, maximum and median $S/N$-ratios per rebinned pixel of the coadded spectra are given in Table~\ref{Tab:log} for the different settings. These values are computed for each spectral order in the reasonably line-free continuum regions used for normalization, and are subsequently applied as flux error estimates.
\section{Analysis}\label{sect:analysis}
\subsection{Effective temperature}\label{subsect:phot}
A photometric effective temperature ($T_{\mbox{{\scriptsize eff}}}$) scale was constructed from the $(v-y)$ and $(b-y)$ colour indices using relations for main sequence stars \citep{Alonso1996,Korn2007} and giant stars \citep{Alonso1999,AlonsoEratum, Korn2007}, assuming $\left[\mbox{Fe/H}\right] =-1.60$ for all stars. Both relations are calibrated on the infrared flux method. As our sample stretches from the TOP at the very end of the MS to the bRGB, we evaluate separately the dwarf and giant calibrations, and interpolate $T_{\mbox{{\scriptsize eff}}}$\ linearly as a function of $\log g$\ between them.
Comparing the $(v-y)$ and $(b-y)$ calibrations, differences are rather small: the former is hotter by $16 \pm 57$\,K for the dwarf and $6 \pm 55$\,K for the giant calibration. The difference is at most roughly 80\,K in a narrow part of the sub-giant branch, which falls outside the colour ranges covered by the calibrations. Temperatures based on $(v-y)$ and $(b-y)$ exhibit a scatter about the fiducial sequences of 40\,K and 80\,K, respectively. Comparing the fiducial sequences on the two colour indices, we find a dispersion of just 23\,K, including the aforementioned biases.
While the $v$-band is in principle sensitive to variations in chemical composition, i.e. different CN band strengths \citep{Carretta2011}, our fiducial sequence does not appear to be strongly affected by this. We judge potential errors in placement of the fiducial sequence due to such effects to be at most $+20$\,K, for our coolest stars where the sensitivity is greatest. We opt to base our temperature scale on the $(v-y)$ relation as it exhibits less scatter, higher sensitivity on the RGB, and creating its fiducial appears to be less susceptible to systematic errors. Taking into account the statistical uncertainties present, we adopt a representative uncertainty of 40\,K in $T_{\mbox{{\scriptsize eff}}}$.
The calibration of \citet{Casagrande2010} on the $(b-y)$ index produces higher temperatures by 120--150\,K, with a slight dependence on colour. No corresponding calibration is available for $(v-y)$. With the main difference to previous calibrations being in the zero-point (as indicated in that paper), we may assume a similar absolute uncertainty would apply to the $(v-y)$ index. \\
\subsection{Surface gravity}\label{subsect:grav}
We employ photometric surface gravities, assuming a distance modulus of 13.13 \citep[latest web update]{Harris1996}. The bolometric corrections of \citet{Alonso1999,AlonsoEratum} are applied to observed $V$-magnitudes, as a function of metallicity and $T_{\mbox{{\scriptsize eff}}}$, to compute luminosities. Stellar masses are interpolated directly as a function of $T_{\mbox{{\scriptsize eff}}}$\ from a 13.5\,Gyr isochrone \citep[from][]{Gruyters2013}, and range from 0.79 to 0.815\,M$_\odot$. The resulting values of $\log g$\ range from 4.0 at $T_{\mbox{{\scriptsize eff}}}$\ $=6050$\,K (TOP) to 3.0 at $T_{\mbox{{\scriptsize eff}}}$\ $=5160$\,K (bRGB). The final stellar parameters, along with observational parameters, are given in Table\,\ref{Tab:SPs}.
The surface gravities are most sensitive to uncertainties in $T_{\mbox{{\scriptsize eff}}}$, where a change of $+100$\,K propagates into $\log g$\ changes of $+0.04$\,dex on the TOP, or $+0.03$\,dex on the bRGB. An effect of $\delta \log g=+0.01$\,dex would require, e.g., an increase in stellar mass of 0.02\,M$_\odot$, or an increase in $V$-magnitude by 0.025\,mag, be it due to observational uncertainties or errors in distance modulus or bolometric correction. \citet{Sbordone2011} examined the effects of chemical composition variations on isochrones and photometry. Concerning isochrones, they found that only extreme changes in nitrogen or helium abundance could produce any discernible effects. \citet{Cassisi2013} expanded on this work and note a composition-driven increase in $T_{\mbox{{\scriptsize eff}}}$\ by $+30$\,K and $\Delta \log(L/L_\odot) \approx 0.017$, where partial cancellation results in an effect $\delta \log g = +0.01$\,dex. On the bRGB on the other hand, $\delta$$T_{\mbox{{\scriptsize eff}}}$\ $\approx 0$\,K, giving an 0.01\,dex net differential effect.
Thanks to the high precision in $\log g$, gravity-sensitive (majority) species such as \ion{Fe}{ii} and \ion{Ti}{ii} are the least susceptible to errors in the primary stellar parameters, as shown in Sect.~\ref{sec:uncertainties}.
\begin{table*}
\caption{Photometry, data quality and stellar parameters.}
\label{Tab:SPs}
\centering
\begin{tabular}{lrc ccc cccc ccccc}
\hline\hline
ID & \ch{RA(J2000)} & $u$ & $v$ & HR9 & HR10 & HR11 & HR15 & Teff & log g & vmic \\
& \ch{Dec(J2000)} & $b$ & $y$ & (S/N) & (S/N) & (S/N) & (S/N) & (K) & (dex) & (km\,s$^{-1}$) \\ \hline
NGC6752-1003 & 19 10 50.351 & 17.423 & 15.974 & 134 & \dots & \dots & \dots &5174 & 3.10 & 1.24 \\
& $-60$ 02 44.166 & 16.575 & 15.479 \\
NGC6752-1013 & 19 10 56.817 & 17.431 & 15.969 & 126 & \dots & \dots & \dots & 5173 & 3.09 & 1.24 \\
& $-60$ 03 06.365 & 16.579 & 15.464 \\
NGC6752-1028 & 19 11 24.854 & 17.445 & 15.985 & 126 & \dots & \dots & \dots & 5174 & 3.10 & 1.24 \\
& $-59$ 57 06.584 & 16.590 & 15.475 \\
NGC6752-1112 & 19 10 35.277 & 17.585 & 16.096 & 103 & 84 & 96 & \dots & 5192 & 3.16 & 1.25 \\
& $-59$ 57 06.024 & 16.692 & 15.615 \\
\hline
\end{tabular}
\tablefoot{
The full table can be retrieved from CDS/Vizier.
}
\end{table*}
\subsection{Spectrum synthesis}
For our heterogeneous sample of spectra of 194 stars, we opted for an automated spectrum analysis rather than the manual line-by-line approach employed for UVES spectra in Paper~I. We use a modified version of the spectrum synthesis code Spectroscopy Made Easy \citep[SME,][]{Valenti1996,Valenti2005}, set to run without user interaction. Additionally, the code has been modified to allow non-LTE (NLTE) line formation, from a grid of precomputed corrections (tabulations of LTE departure coefficients). To run the code, one provides an input spectrum, stellar parameters ($T_{\mbox{{\scriptsize eff}}}$, $\log g$, $\left[\mbox{Fe/H}\right]$ and $v_{\mbox{{\scriptsize mic}}}$), a line list with line parameters, line and continuum masks, and a list of spectrum segments wherein the continuum is individually normalized. The code uses a grid of MARCS plane-parallel and spherically symmetric model atmospheres \citep{Gustafsson2008}, all with scaled solar abundances and alpha-enhancement of 0.4\,dex. Interpolating models from this grid, spectra are then computed on the fly. The numerical comparison between synthetic and observed spectra is executed by a non-linear optimization algorithm \citep{Marquardt1963,Press1992}. We apply NLTE corrections for Fe, Na and Li \citep{Lind2009, Lind2011,Lind2012} to the line formation using pre-computed departure coefficients. Finally, we apply NLTE corrections to our LTE abundances of Mg (Y. Osorio, priv. comm.) as in Paper~I. \\
\subsection{Microturbulence}
As a large fraction of our stellar sample only has spectra of moderate quality in the rather line-poor spectral region covered in HR15B, $v_{\mbox{{\scriptsize mic}}}$\ cannot be determined to acceptable precision for each star. We executed preliminary analyses where $v_{\mbox{{\scriptsize mic}}}$\ was determined from a set of iron lines in the HR9B setting, and found an essentially linear dependence on effective temperature with an RMS error of 0.11\, km\,s$^{-1}$. Applying this linear relation to all stars, $v_{\mbox{{\scriptsize mic}}}$\ values range between 1.8\,km\,s$^{-1}$\ at $T_{\mbox{{\scriptsize eff}}}$\ $=6000$\,K and 1.3\,km\,s$^{-1}$\ at $T_{\mbox{{\scriptsize eff}}}$\ $=5200$\,K. On our photometric temperature scales, the excitation equilibrium is not necessarily fulfilled. This could skew the determination of $v_{\mbox{{\scriptsize mic}}}$\ due to the tendency for low-$\chi$ lines to be stronger, thus correlating $T_{\mbox{{\scriptsize eff}}}$\ with $v_{\mbox{{\scriptsize mic}}}$. We executed test runs where $T_{\mbox{{\scriptsize eff}}}$\ was determined simultaneously with $v_{\mbox{{\scriptsize mic}}}$\ from \ion{Fe}i lines, resulting in overall significantly higher temperatures, as well as a somewhat expanded temperature scale. The difference is $\delta $$T_{\mbox{{\scriptsize eff}}}$\ $= 207 \pm 143$\,K and $60\pm48$\,K for the warm ($T_{\mbox{{\scriptsize eff}}}$\ $> 5800$\,K) and cool ($T_{\mbox{{\scriptsize eff}}}$\ $< 5450$\,K) stars of our sample, with corresponding increases $\delta$$v_{\mbox{{\scriptsize mic}}}$\ $=0.10 \pm 0.16$ and $0.08\pm0.07$\,km\,s$^{-1}$, respectively. Deriving $v_{\mbox{{\scriptsize mic}}}$\ from \ion{Fe}{ii} lines, as was done in Paper I, we find only a minor difference $\delta$$v_{\mbox{{\scriptsize mic}}}$\ $=-0.06 \pm 0.20$\,km\,s$^{-1}$\ for the warm, and $\delta$$v_{\mbox{{\scriptsize mic}}}$\ $=0.01\pm0.06$\,km\,s$^{-1}$\ for the cool stars. With these considerations in mind, we adopt 0.10\,km\,s$^{-1}$\ as our representative uncertainty. At this level, the $v_{\mbox{{\scriptsize mic}}}$\ uncertainty does not add significantly to the error budget of derived abundances (cf. Sect.~\ref{sec:uncertainties}).
\subsection{Deriving chemical abundances}
We derive chemical abundances for each star independently from each available spectral order. This allows us to check the order-to-order consistency, and ensures internal homogeneity among abundances derived from a single spectral order. Only after ensuring consistent results do we average results from different spectral orders, where relevant. As spectral orders vary in the amount and quality of spectral information, they are weighted by statistical uncertainties in the averaging.
To further maintain consistency between settings and reduce (internal) line-to-line scatter, we determine abundances differentially to a high-quality averaged spectrum on the cool end of our sample for the $\alpha$ and iron-peak elements. We selected all 11 stars on the bRGB ($T_{\mbox{{\scriptsize eff}}}$\ $< 5420$\,K) with complete sets of observations in all spectral orders. By coadding these spectra, we achieve S/N well above 200 in all settings. We executed a preliminary abundance analysis on these spectra, and adopted these results as our reference abundance pattern. Line oscillator strengths were then adjusted to reproduce this abundance pattern, so that the baseline (at the bRGB) is the same for all lines, in all spectral orders. The reference stars are detailed in Table~\ref{Tab:SN-coadd}.
\begin{table}
\caption{Coadded bRGB spectra used as reference in differential analysis.}
\label{Tab:SN-coadd}
\centering
\begin{tabular}{lcccccccc}
\hline\hline
Star ID & S/N & S/N & S/N & S/N & $T_{\mbox{{\scriptsize eff}}}$ & $\log g$ \\
& HR9 & HR10 & HR11 & HR15 & (K) & (dex) \\ \hline
1872 & 89 & 72 & 64 & 63& 5302 & 3.46 \\
1980 & 76 & 69 & 90 & 132& 5309 & 3.48 \\
2113 & 73 & 63 & 63 & 58& 5328 & 3.51 \\
2089 & 80 & 65 & 68 & 57& 5328 & 3.51 \\
2129 & 73 & 69 & 70 & 59& 5332 & 3.52 \\
2176 & 71 & 61 & 97 & 135& 5339 & 3.54 \\
201217 & 56 & 55 & 80 & 114& 5355 & 3.57 \\
2334 & 69 & 58 & 74 & 106& 5355 & 3.57 \\
2396 & 76 & 57 & 87 & 117& 5358 & 3.57 \\
2337 & 68 & 61 & 90 & 123& 5361 & 3.57 \\
2646 & 75 & 67 & 86 & 106& 5413 & 3.62 \\\hline
coadded & 224 & 212 & 267 & 259 & 5344 & 3.54 \\
\hline\hline
\end{tabular}
\end{table}
When the chemical composition exhibits internal variations among the coadded stars, nontrivial problems may arise from the averaging of lines whose strengths and shapes are intrinsically different. For this reason, we do not adopt astrophysical oscillator strengths for the light elements where such variations are expected: Li, Na, Mg, and Al. Note however that most of these elemental abundances are derived from only one or two lines, and thus would anyway not be affected by this issue. The atomic data for these lines can be found in Table~\ref{Tab:atomic data}. The other atomic data for the lines used in the abundance analysis were obtained from the Vienna Atomic Line Database \citep{piskunov1995,Kupka1999}.
\begin{table}
\caption{Atomic Line Data.}
\label{Tab:atomic data}
\centering
\include{linelist}
\end{table}
\subsection{Uncertainties} \label{sec:uncertainties}
We consider systematic and statistical errors separately.
Leading systematic error sources likely stem from uncertainties in the atmospheric parameters, which were discussed in Sect.~\ref{subsect:phot} and \ref{subsect:grav}. Systematic abundance uncertainties from the influence of stellar parameter perturbations on average abundance results are indicated in Table~\ref{Tab:abund sensitivity}. We evaluate these effects for the hot and cool ends of our temperature range, for all chemical elements analysed in this work.
Modelling errors related to the use of 1D atmospheres were discussed in Paper~I.
Lines of the same ionisation stage respond, in general, in a similar way to changes in atmospheric parameters. Thus the systematic uncertainties related to the atmospheric parameters partially cancel when working with abundance ratios ($\left[\mbox{X/Y}\right] = [{\rm X}/{\rm H}] - [{\rm Y}/{\rm H}]$).
For statistical errors, we use those reported by the $\chi^2$ minimization algorithm of SME \citep{Marquardt1963,Press1992}, which are based on photon-noise statistics. We shall refer to these as 'SME errors'. As they do not fully take into account uncertainties in the continuum determination, we enforce a representative minimum error of 0.02\,dex. Unless otherwise stated, uncertainties in this work refer to these statistical errors.
We examine the estimate of these statistical errors in more detail by running a Monte Carlo simulation, in which we derive the abundance from a single weak atomic line, representing the case of lithium. As reference, we use synthetic spectra generated for stellar parameters ($T_{\mbox{{\scriptsize eff}}}$\ $=5400$--6000\,K) and abundances spanning the ranges of the stars observed ($EW = 7$--45\,m\AA), with a wide range of Gaussian noise injected ($S/N = 10$--1000). We generated a total of 65 such sets of at least 10\,000 spectra each. We averaged results from each set of abundance analyses, taking SME errors as uncertainties. The weighted standard deviation and the bias of the mean from these averages are compared in Fig.~\ref{fig:errors} to the SME error. Small errors appear to be dominated by the width of the distribution, such that averaging a large sample will generate an accurate estimate of the mean. Large errors on the other hand tend to become asymmetric, giving a significant positive bias, while underestimating the true dispersion. The figure can be interpreted as a decomposition of the SME error into these two components, for any single abundance result. Or, the bias of an averaged abundance can be estimated from the size of its standard deviation. For instance, averaging a large sample of lithium abundances with 0.10\,dex uncertainties would overestimate the mean value by 0.03\,dex.
The cause of this behaviour is that noise produces errors in line equivalent widths which are symmetric and linear. For weak lines, the same is then true for the number of absorbers $N_{\rm X}/N_{\rm H}$. On the logarithmic abundance scale, $(\log N_{\rm X}/N_{\rm H})$, these errors instead appear asymmetric, such that weighted averages always systematically overestimate the mean value. If instead lines are on the saturated or strong parts of the curve of growth, the corresponding error statistics apply. Finally, these considerations would not necessarily apply if uncertainties in stellar parameters dominate over those in line strengths.
As we analyze lines on various parts of the curve of growth, we choose not to apply linear averages. Tests indicate that the magnitude of these biases, if applicable, is less than 0.03\,dex in this work.
\begin{figure}
\includegraphics[]{mc_errors}
\caption{Weighted standard deviations and biases of the weighted means as determined from weak lines analysed in Monte Carlo tests on synthetic spectra.}
\label{fig:errors}
\end{figure}
\begin{table}
\caption{Abundance sensitivity to stellar parameters.}
\label{Tab:abund sensitivity}
\centering
\begin{tabular}{l r r r r r r}\hline\hline \noalign{\smallskip}
Ion & \multicolumn 2 c {$T_{\mbox{{\scriptsize eff}}} + 100$\,K} & \multicolumn 2 c {$\log g + 0.1$\,dex} & \multicolumn 2 c {$v_{\mbox{{\scriptsize mic}}} + 0.1$\,km\,s$^{-1}$} \\
& \ch{TOP} & \ch{bRGB} & \ch{TOP} & \ch{bRGB} & \ch{TOP} & \ch{bRGB} \\ \hline\noalign{\smallskip}
Li I & $ 0.073$ & $ 0.078$ & $-0.002$ & $-0.002$ & $-0.001$ & $ 0.006$ \\
Na I & $ 0.032$ & $ 0.046$ & $ 0.002$ & $ 0.000$ & $ 0.000$ & $-0.001$ \\
Mg I & $ 0.080$ & $ 0.116$ & $-0.030$ & $-0.050$ & $-0.014$ & $-0.007$ \\
Al I & \dots\hphantom{..} & $ 0.037$ & \dots\hphantom{..} & $ 0.001$ & \dots\hphantom{..} & $ 0.004$ \\
Si I & $ 0.025$ & $ 0.036$ & $ 0.004$ & $ 0.008$ & $ 0.009$ & $-0.001$ \\
Ca I & $ 0.048$ & $ 0.065$ & $ 0.000$ & $-0.002$ & $-0.007$ & $-0.015$ \\
Sc II & \dots\hphantom{..} & $ 0.043$ & \dots\hphantom{..} & $ 0.035$ & \dots\hphantom{..} & $-0.006$ \\
Ti II & $ 0.038$ & $ 0.050$ & $ 0.032$ & $ 0.026$ & $-0.010$ & $-0.025$ \\
Cr I & \dots\hphantom{..} & $ 0.132$ & \dots\hphantom{..} & $-0.020$ & \dots\hphantom{..} & $-0.038$ \\
Fe I & $ 0.075$ & $ 0.109$ & $-0.006$ & $-0.014$ & $-0.014$ & $-0.027$ \\
Fe II & $ 0.040$ & $ 0.058$ & $ 0.022$ & $ 0.013$ & $-0.019$ & $-0.031$ \\
Ni I & \dots\hphantom{..} & $ 0.073$ & \dots\hphantom{..} & $ 0.004$ & \dots\hphantom{..} & $-0.004$ \\
\hline\hline
\end{tabular}
\tablefoot{Effects on the averaged abundances are shown for the hot and cool ends of our sample, as defined in Table~\ref{Tab:mean-trends}.
}
\end{table}
\section{Results}
In what follows we address our derived chemical abundances for the total sample of 194 stars. As not all stars were observed with all settings, we do not have a complete set of abundances for the full sample. Results from different spectral orders have been averaged only after carefully verifying that they appear compatible. After describing our abundance results, we discuss the implications, and investigate the prospect of predicting population membership from photometry.
\subsection{Chemical abundances}
\subsubsection{Lithium} \label{sec:lithium}
Lithium abundances $A$(Li) are given in Table~\ref{Tab:pop-abun} and are derived from the \element[][7]{Li} resonance line at 6707.8\,\AA. The line has two fine-structure components which are separated by only 0.15\,\AA\ and thus are indistinguishable at the resolution of GIRAFFE ($R = 19\,300$). The atomic data for the Li doublet takes into account this fine structure and isotopic splitting and is given in Table~\ref{Tab:atomic data}. The abundances are corrected for NLTE effects by interpolation on the grid by \citet{Lind2009}. As lithium is almost completely ionised, $A$(Li) is primarily sensitive to $T_{\mbox{{\scriptsize eff}}}$, which thus dominates the systematic errors. Our estimated precision of 40\,K in $T_{\mbox{{\scriptsize eff}}}$\ corresponds to an abundance uncertainty of 0.03~dex. Compared to this, errors due to uncertainties in gravity, metallicity and microturbulence are generally negligible ($\sim 0.01$\,dex).
Combining these systematic uncertainties, with the expected bias stemming from the averaging process, we estimate 0.10\,dex to be a suitable conservative error. The average Li abundance of the TOP sample ($T_{\mbox{{\scriptsize eff}}}$ > 5900\,K) is $2.22\pm0.08$ in agreement with Paper~I and \citet{Gratton2001}. We will discuss lithium in greater depth in Sect.~\ref{sec:evol-lithium}.
\subsubsection{$\alpha$ and iron-peak elements} \label{sec:aip}
Abundances of the $\alpha$ elements magnesium, calcium and titanium were derived along with the iron-peak elements scandium, chromium, iron and nickel.
\begin{figure*}
\begin{center}
\includegraphics[]{abund_evo.eps}
\caption{Evolutionary abundance trends of Mg, Ca, Ti, and Fe. Mg and Ca abundances are derived from neutral lines while Ti and Fe are derived from singly ionised lines (see Table~\ref{Tab:mean-trends}). The solid (green) line represents the running mean (weighted average), while dashed (blue) lines represent the standard deviation. Error bars represent statistical errors. Systematic errors are indicated in Fig.~\ref{Fig:AD-trends}.
Overplotted in diamonds are the UVES results from Paper~I, which have been arbitrarily shifted for easier comparison.}
\label{Fig:abundances}
\end{center}
\end{figure*}
Figure~\ref{Fig:abundances} shows the derived chemical abundances of magnesium, calcium, titanium and iron as a function of the effective temperature. Results have been averaged on a running mean in bins of $\pm$150\,K, which has then been further smoothed by a Gaussian ($\sigma = 100$\,K).
We show also the group-averaged abundances derived from UVES spectra in Paper~I, from a temperature scale based on the $(b-y)$ colour index.
We have shifted those results slightly to coincide with our abundance scale for the coolest stars.
Magnesium abundances were derived from the $\lambda\lambda$5167--5184 triplet in the HR9 setting and the high-excitation $\lambda$5528 line in the HR10 setting.
We average abundances between the two settings for the 50 stars which were observed in both settings, as we find rather consistent results when adopting laboratory oscillator strengths \citep{Chang1990,Aldenius2007} and applying NLTE corrections. The rather large observed dispersion, compared to the other elements, is further discussed in Sect.~\ref{sec:anticorr}.
Calcium abundances were likewise determined from the HR9, HR10 and HR11 settings. When applicable, we average results between settings, resulting in abundances determined for a set of 176 stars. We investigated departures from LTE, following \citet{Mashonkina2007}, for several of the \ion{Ca} i lines used in the determination of the Ca abundance. Apart from the general effect of over-ionisation which raises all abundances by $\sim 0.1$\,dex, the relative trends between TOP and RGB are only marginally affected, on average by 0.02\,dex. This correction is well within the error budget of our analysis.
Titanium and iron abundances have been derived from ionized lines, which we selected exclusively from HR9. As only half the stars were observed with this setting, we also determined iron abundances for each star from the neutral lines available in every spectral order. These abundances were corrected for NLTE effects and averaged between settings, but are only used as a reference in computing $\left[\mbox{X/Fe}\right]$.
\begin{table*}
\caption{Average abundances obtained at two effective temperature points.}
\label{Tab:mean-trends}
\centering
\begin{tabular}{lcccrrrr}\hline\hline
Group & $T_{\mbox{\scriptsize eff}}$ & $\log g$ & $\xi$ & \ch{$\abund{Mg}$\tablefootmark{a}} & \ch{$\abund{Ca}$\tablefootmark{b}} & \ch{$\abund{Ti}$\tablefootmark{c}} & \ch{$\abund{Fe}$\tablefootmark{d}} \\
& (K) & (cgs) & (km\,s$^{-1}$) & \ch{NLTE} & \ch{LTE} & \ch{LTE} & \ch{NLTE} \\ \hline
TOP & 6050 & 4.02 & 1.8 & $ 6.02 \pm0.08$ & $ 4.88 \pm0.05$ & $ 3.53 \pm0.05$ & $ 5.75 \pm0.04$ \\
bRGB & 5160 & 3.03 & 1.2 & $ 6.13 \pm0.09$ & $ 4.98 \pm0.04$ & $ 3.62 \pm0.04$ & $ 5.83 \pm0.03$ \\
$\Delta(\text{TOP}-\text{bRGB})$ & 890 & 0.99 & 0.6 & $-0.11 \pm0.12$ & $-0.11 \pm0.06$ & $-0.09 \pm0.06$ & $-0.08 \pm0.05$ \\
\hline\hline
\end{tabular}
\tablefoot{
\tablefoottext{a}{Based on Mg\,{\scriptsize I} $\lambda\lambda$5167, 5173, 5184 and 5528.}
\tablefoottext{b}{Based on Ca\,{\scriptsize I} $\lambda\lambda$5261, 5262, 5265, 5349, 5582, 5589, 5590, 5594, 5598, 5601 and 5603.}
\tablefoottext{c}{Based on Ti\,{\scriptsize II} $\lambda\lambda$5154, 5185, 5188, 5226 and 5336.}
\tablefoottext{d}{Based on Fe\,{\scriptsize II} lines $\lambda\lambda$5146, 5169, 5197, 5234, 5264, 5276, 5284 and 5316.}}
\end{table*}
All elements exhibit clear evolutionary variations, summarized for the hot and cool ends of our temperature range in Table~\ref{Tab:mean-trends}, where we also state explicitly which spectral lines have been used.
Although these abundance variations are rather weak and of low significance (1--$2\sigma$), they are in good agreement with Paper~I, despite the independent methods used here. The abundance variations are also rather robust regarding systematic errors in stellar parameters. For instance, removing the abundance trend in iron would require increasing temperatures of stars at the TOP by 200\,K, implying an error in $\Delta\text{$T_{\mbox{{\scriptsize eff}}}$}(\text{TOP}-\text{bRGB})$ of 25\,\%, which we consider unlikely. A similarly large adjustment to $\log g$\, (+0.35\,dex) or $v_{\mbox{{\scriptsize mic}}}$\ (--0.4\,km\,s$^{-1}$) would be required. Systematic uncertainties of stellar parameters were further examined by \citet{Nordlander2012}, indicating that no combination of stellar parameters can simultaneously render null trends in all elements. Errors due to hydrodynamical effects were investigated in Paper~I. A full 3D treatment (in LTE) would in fact strengthen all observed trends.
\begin{table}
\caption{Average abundances and dispersions of Sc, Cr, Ni and Fe for the subsample of bRGB stars ($T_{\mbox{{\scriptsize eff}}}$\ $<5450$\,K).}
\label{Tab:abundances}
\centering
\begin{tabular}{lcccc}\hline\hline
& Sc\tablefootmark{a} & Cr\tablefootmark{b} & Fe\tablefootmark{c} & Ni\tablefootmark{d}\\
& LTE & LTE & NLTE & LTE \\ \hline
$\log\varepsilon$ & 1.52 & 3.88 & 5.83 & 4.48 \\
$\sigma(\log\varepsilon)$ & 0.05 & 0.05 & 0.03 & 0.03 \\
\hline\hline
\end{tabular}
\tablefoot{The dispersion refers to the weighted standard deviation around the mean.
\tablefoottext{a}{Based on Sc\,{\scriptsize II} $\lambda\lambda$5527 and 5658.}
\tablefoottext{b}{Based on Cr\,{\scriptsize I} $\lambda\lambda$5206--5208.}
\tablefoottext{c}{Based on Fe\,{\scriptsize II} lines $\lambda\lambda$5146, 5169, 5197, 5234, 5264, 5276, 5284 and 5316.}
\tablefoottext{d}{Based on Ni\,{\scriptsize I} $\lambda\lambda$5146, 5155, 5477, 6643 and 6768.}
}
\end{table}
We also derived abundances of scandium, chromium and nickel for the bRGB stars in our sample. Unfortunately, the quality of the dwarf star spectra did not allow to derive reliable abundances for these elements in these stars. An overview of the mean abundances and dispersion for the bRGB stars ($T_{\mbox{{\scriptsize eff}}}$\ $< 5450$\,K) is given in Table~\ref{Tab:abundances}. \\
The abundances of Ca, Ti, Cr, Fe and Ni are in agreement with the abundances derived for the reference RGB bump star NGC6752--11 from the \citet{Yong2013} sample. Unfortunately, they did not derive Mg or Sc abundances. The abundances show the typical pattern of halo field stars in which the $\alpha$-elements are overabundant by 0.1 to 0.3\,dex relative to the Fe-peak elements which are roughly solar, meaning $\left[\alpha\mbox{/Fe}\right]\approx0.1$--0.3 and $\left[\mbox{X/Fe}\right]\approx0$. \\
Three of the stars examined here (IDs 2844, 2795, 3035) exhibit suspiciously low abundances in calcium, titanium and iron, which could indicate binarity. For instance, an evolved low-mass companion would only moderately affect photometric parameters, but rather provide continuum flux filling in all spectral lines.
\begin{figure*}
\begin{center}
\includegraphics[]{anticorr_evo.eps}
\caption{Evolutionary abundance trends of Na, Mg, Al, and Si. The different coloured symbols correspond to the chemical populations deduced from a photometric cluster analysis (see text).}\label{Fig:pop_abun}
\end{center}
\end{figure*}
\subsubsection{Na, Mg, Al and Si} \label{sec:anticorr}
Distinct chemical populations in globular clusters have previously been addressed in a series of papers (on NGC\,6752, e.g., by \citealt{Gratton2001} and \citealt{Carretta2007a, Carretta2012}).
In Paper~I, we did the same for a set of FLAMES-UVES spectra, revealing three chemically distinct populations in agreement with \citet{Carretta2012}. Based on the abundances of Al, Na and Mg we labelled stars to be first- or second-generation stars. First-generation stars are poor in Al and Na while rich in Mg, compared to the mean of the whole sample. This group of stars is referred to as the primordial population (P). Second-generation stars are characterised by average and high Al and Na abundances compared to the mean abundances for our sample. This second generation can be further split into two different populations: one with average abundances of Al, Na and Mg (intermediate population I), and one rich in Al and Na but poor in Mg (extreme population E). Si is not considered a discriminating element in this sense. \\
Figure~\ref{Fig:pop_abun} shows the abundances of sodium, magnesium, aluminium and silicon as a function of $T_{\mbox{{\scriptsize eff}}}$. All elements exhibit a large dispersion, as expected from previous findings (e.g. \citealt{Carretta2007a,Carretta2012}; Paper~I).
Symbols and colours in this figure correspond to the different chemical populations, determined photometrically in Sect.~\ref{sec:clusteranalysis}.
Sodium abundances were derived for a set of 132 stars using the $\lambda\lambda$5682--5688 doublet in HR11. We find a spread (tip-to-tip) of roughly 0.8\,dex, which does not appear to vary significantly along the sequence, sans sampling effects. While \citet{Lind2011} found a 0.2\,dex difference between dwarf and giant stars in the cluster NGC\,6397, we do not observe a clear difference in abundance between the two groups. \citet{Lind2011} explain their observed difference as a result from atomic diffusion. At the present metallicity, the expected trend is shallower ($\sim 0.1$\,dex), while the star-to-star dispersion is similar. A potential trend is thus difficult to detect using the very weak lines available in our spectra. We will address atomic diffusion in more detail in Sect.~\ref{subsect:AD}.
By contrast, our magnesium abundances exhibit a smaller spread of just 0.3\,dex. The spread again does not exhibit significant variation along the sequence. The abundance difference comparing warm to cool stars, $0.11\pm0.12$\,dex (Table~\ref{Tab:mean-trends}), is similar for the three populations, $0.16 \pm 0.06$, $0.16 \pm 0.12$ and $0.12 \pm 0.14$\,dex, for the P, I and E populations, respectively. As the P and I populations exhibit similar abundances of magnesium, we note that their combined sample indicates an abundance trend of $0.15 \pm 0.08$\,dex.
Aluminium abundances are derived from the weak \ion{Al} i doublet at 6696--6698\,\AA\ and thus could only be derived from stars observed in HR15. Unfortunately, the doublet was too weak to be observed in most of the stars, and even upper limits were not meaningful for the bulk of (lower-S/N) spectra.
In total, we could detect aluminium in 22 stars, and derived upper limits for another 48, indicating a spread of at least 1\,dex.
Finally, we deduced silicon abundances from three Si\,{\scriptsize I} lines in HR11B at 5708, 5665 and 5690\,\AA. The spread in Si abundance is dominated by the observational uncertainty on the hotter end of our sample. This uncertainty also masks out any potential trend with evolutionary phase. We estimate an intrinsic spread in Si of about 0.3\,dex.
The abundances ratios of Na, Mg, Al and Si relative to Fe (based on \ion{Fe} i lines) for the individual stars are given in Table~\ref{Tab:pop-abun}, and compared in Fig.~\ref{Fig:AbunRatio}. Using Mg as the discriminator, the usual anticorrelations are observed, a signature of element depletion in H burning at high temperature. We also see that Al is well correlated with elements which are enhanced by the Ne-Na and Mg-Al cycles. And like \citet{Carretta2012}, we also seem to find a Al-Si correlation. The correlation stems from leakage from the Mg-Al cycle on $^{28}$Si. Such a leakage can only be obtained within a reaction network in which the temperature exceeds $\sim 65 \times 10^6$\,K. The Al-Si correlation can thus be explained as a direct result of pollution by the first generation of cluster stars burning H at high temperatures.
\begin{table*}
\caption{Light element abundances and abundance ratios, compared to photometric population assignments.}
\label{Tab:pop-abun}
\centering
\begin{tabular}{lccccccc}
\hline\hline
Star ID & $A$(Li) & $\left[\mbox{Na/Fe}\right]$ & $\left[\mbox{Mg/Fe}\right]$ & $\left[\mbox{Al/Fe}\right]$ & $\left[\mbox{Si/Fe}\right]$ & $\log\varepsilon$(Fe) & Pop\tablefootmark{a} \\
& NLTE & NLTE & NLTE & LTE & LTE & NLTE & \\ \hline
1802 & $ <0.82$ & \dots & \dots & $ 0.78 \pm 0.08$ & \dots & $ 5.81 \pm 0.03$ & E \\
1980 & $ 1.31 \pm 0.04$ & $ 0.10 \pm 0.04$ & $ 0.27 \pm 0.03$ & $ 0.49 \pm 0.05$ & $ 0.26 \pm 0.04$ & $ 5.81 \pm 0.02$ & I \\
1983 & $ <1.05 $ & $ 0.50 \pm 0.03$ & \dots & $ 0.63 \pm 0.05$ & $ 0.36 \pm 0.05$ & $ 5.79 \pm 0.02$ & E \\
2117 & $ 1.19 \pm 0.07$ & $ 0.03 \pm 0.04$ & $ 0.33 \pm 0.03$ & $ <0.42 $ & $ 0.21 \pm 0.05$ & $ 5.76 \pm 0.02$ & P \\
2129 & $ 1.34 \pm 0.09$ & $ 0.27 \pm 0.04$ & $ 0.22 \pm 0.03$ & $ <0.29 $ & $ 0.25 \pm 0.06$ & $ 5.81 \pm 0.02$ & I \\
2151 & $ 1.00 \pm 0.07$ & $ 0.31 \pm 0.04$ & \dots & $ 0.58 \pm 0.05$ & $ 0.32 \pm 0.06$ & $ 5.76 \pm 0.02$ & I \\
\hline\hline
\end{tabular}
\tablefoot{Uncertainties are given by statistical errors. All abundances are based on lines of the neutral species. Upper limits are given by $<$.\\
\tablefoottext{a}{Photometrically identified chemical population: primordial (P), intermediate (I) and extreme (E).}\\ The full table can be retrieved from CDS/Vizier. \\
Solar reference abundances: $A(\text{Na}) = 6.17$, $A(\text{Mg}) = 7.53$, $A(\text{Al}) = 6.37$, $A(\text{Si}) = 7.51$, $A(\text{Fe}) = 7.45$.
}
\end{table*}
\begin{figure*}
\begin{center}
\includegraphics[]{anticorr_cf.eps}
\caption{Anticorrelations between sodium, magnesium and aluminium. Different coloured symbols correspond to the chemical populations deduced from a photometric cluster analysis (see text). The faintest ($V > 16.8$) stars have been excluded from this comparison.}\label{Fig:AbunRatio}
\end{center}
\end{figure*}
In the next section we will address these distinct chemical populations in greater depth by investigating their photometric properties.
\subsection{A cluster analysis}\label{sec:clusteranalysis}
It has previously been demonstrated by \citet{Carretta2012} that one can couple abundance anticorrelations to photometric properties of RGB stars in NGC\,6752. We shall here attempt to extend their findings for fainter stars, along the SGB toward the TOP region.
Using $uvby$ Str\"omgren photometry for our complete data set, we created colour-magnitude fiducial sequences for the colour indices $(u-v)$, $(u-b)$, $c_y$ and $\delta_4$. Figure~\ref{Fig:CMD-abun} shows the corresponding CMDs with the fiducial sequences overplotted. The $c_y$ index was defined by \citet{Yong2008} and combines three colour indices: $c_y = (u-v) - (v-b) - (b-y) = u - 2\times v + y$. It is sensitive to N abundances, from the influence on NH bands near 3450\,\AA\ in the $u$ filter, and CN bands at 4216\,\AA\ in the $v$ filter. As the sensitivity on the latter is weaker, it is weighted twice. The $\delta_4 = (u-v) - (b-y)$ index was introduced by \citet{Carretta2011} and combines the four Str\"omgren filters, each weighted only once. For further in-depth discussion of the $c_y$ and $\delta_4$ indices, the reader is referred to \citet{Carretta2011}. \\
We now pose the following question: Based solely on photometry, can we predict membership of our globular cluster stars to different chemical populations? It was shown by \citet{Sbordone2011} and \citet{Cassisi2013} that the effect of anticorrelated Mg and Al variations on Johnson and Str\"omgren photometry is negligible. However, these anticorrelations appear intrinsically tied to variations in CNO and He, which affect the spectral energy distribution (SED) both directly via molecular band strengths, and indirectly via secondary effects on the stellar evolution. We here make an attempt to predict chemical population membership based on photometric information alone. In a next step we then compare these predictions to the derived abundances for Na, Mg, Al and Si. This was done as follows.
With the photometric indices described above, we performed a cluster analysis using the $k$-means algorithm \citep{Steinhaus1956,MacQueen1967} as implemented in the $R$ statistical package \citep{R2013}\footnote{http://www.R-project.org}. We consider only photometric parameters for each star and take the residuals of colour indices with respect to the fiducial sequences at constant $V$ magnitude as the parameters for the cluster analysis. We do not take into account the varying widths of the distributions as a normalizing factor. By the nature of the $k$-means method, uncertainties are not taken into account, and populations are assigned as definitive rather than probabilistic. The resulting average values of the photometric parameters for stars in three populations as determined by our cluster analysis are given in Table~\ref{Tab:clusters}. \\
\begin{table}
\caption{Average photometric parameters of cluster populations}
\label{Tab:clusters}
\centering
\begin{tabular}{lccccc}
\hline\hline
Pop. & nr & $\Delta u-v$ & $\Delta u-b$ & $\Delta c_y$ & $\Delta \delta_4$ \\ \hline
P & 46 & -0.030 & -0.027 & -0.033 & -0.033 \\
I & 101 & 0.001 & 0.001 & -0.001 & 0.001 \\
E & 47 & 0.026 & 0.019 & 0.031 & 0.035 \\
\hline\hline
\end{tabular}
\tablefoot{Stars have been identified by a $k$-means analysis on photometric data.}
\end{table}
With all stars in our sample assigned to populations, we return to the light element abundances. In Fig.~\ref{Fig:AbunRatio} we show the relations between abundances of sodium, magnesium and aluminium, colour-coded by population. We find stars low in sodium and aluminium but rich in magnesium to be generally correctly identified as first-generation stars. Stars rich in sodium and aluminium but low in magnesium are labeled as extreme stars belonging to the second generation. There is, however, some crosstalk between the primordial and intermediate population on the one hand, and between the extreme and intermediate populations on the other.
Since many of our abundance results for these elements are only upper limits, it is difficult to say definitively whether stars have truly been mislabeled. From Fig.~\ref{Fig:CMD-abun}, more crosstalk is seen among the faintest ($V > 16.8$), warmest ($T_\text{eff} > 5900$\,K) stars of the sample. This should be expected due to weaker correlation between photometry and chemistry on the merit of weaker CN-bands. Spectroscopic assignments are unfortunately similarly difficult as a result of limited $S/N$ for these faint stars, making it difficult to derive precise abundances from weak lines (see Fig.~2 in Paper~I).
Comparing the statistics of our population assignments to those by \citet{Carretta2012}, we find overall similar results.
The primordial population comes out the smallest at 24\,\%, at a similar fraction to the 21\,\% of \citet{Carretta2012} and 25\,\% of \citet{Milone2013}. A general fraction of $\sim 30$\,\% primordial stars in GCs was determined by \citet{Carretta2013} by comparing the Na abundances of field stars with GC stars.
We find a twice larger intermediate than extreme population, while \citet{Carretta2012} find the two to be roughly comparable, and \citet{Milone2013} find a 3:2 ratio. Our result may be somewhat inflated due to shortcomings in our analysis relating to the weaker photometric sensitivity to chemical composition toward the TOP. Firstly, the photometric variations are simply weaker in relation to the observational uncertainty. Secondly, the fact that the variations weaken toward higher temperature leads to any warm star regardless of chemical composition appearing more similar to an intermediate (cool) star.
We have also directly compared to \citet{Carretta2012} by including their stars in the cluster analysis. Crossmatching our photometric catalogue to their sample of stars, the three populations are identified in the same proportions, to within 1\,\%. Among the stars they assigned to the P, I and E populations, we misidentify 20, 31 and 17\,\%, respectively. Most mismatches are between the I and E populations, while only one star in the E population is assigned P membership.\\
We conclude that a cluster analysis based solely on $uvby$ Str\"omgren photometry can indeed separate GC stars into first- and second-generation populations on the subgiant branch, with some success even down toward the TOP. We explore this possibility further in Sect.~\ref{sec:evol-lithium}. As expected though, significant crosstalk arises when the photometric response to chemical composition weakens to a level comparable to photometric uncertainties. Finally, we would like to note that the separation of GC stars in populations based on $uvby$ Str\"omgren photometry is the direct result of abundance variations in C, N and O, and in He. However, since O correlates with Mg but anticorrelates with N, stars rich in O/Mg will be poor in N and vice versa. This variation in N abundance is then responsible for the separation in colour index upon which we base our method of identifying chemical populations. Besides the effect of varying N abundances in the different populations, \citet{Milone2013} also discussed the influence of different He fractions, $Y$. However, as we have no direct information on the He content of our stars, we can only refer to their results that the first generation of stars is characterised by a primordial He abundance $(Y\sim0.246)$, is rich in C/O/Mg and poor in N/Na/Al. The extreme population is made up of second-generation stars enhanced in He $(\Delta Y\sim0.03)$ and N/Na/Al but depleted in C/O/Mg. Finally the intermediate population also represents second-generation stars, but with intermediate He $(\Delta Y\sim0.01)$, C/O/Mg and N/Na/Al abundances.
\begin{figure}
\begin{center}
\includegraphics[]{CMDs_small.eps}
\caption{CMDs in $(u-v)$ (upper left), $(u-b)$ (upper right), $\delta_4$ (lower left), and $c_y$ (lower right). The solid lines represent the fiducial sequences from which distances (residuals) are used in the cluster analysis. Different coloured symbols correspond to the chemical populations deduced from a cluster analysis.}\label{Fig:CMD-abun}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[]{uvby_selection_fullphot.eps}
\caption{CMDs in $(u-v)$ (upper left), $(u-b)$ (upper right), $\delta_4$ (lower left), and $c_y$ (lower right). The solid (blue) lines represent the fiducial sequences from which distances (residuals) are used in the cluster analysis.
Blue diamonds indicate points between which the sequence has been interpolated.
Yellow squares represent the selected stars for the cluster analysis.}\label{Fig:phot selection}
\end{center}
\end{figure}
\begin{figure*}
\begin{center}
\includegraphics[]{uvby_populations_fullphot.eps}
\includegraphics[]{uvby_location.eps}
\caption{{\sl Left}: Cumulative radial distribution of the different populations as a function of distance from the cluster center, {\sl Right}: Spatial distribution of the stars in NGC\,6752. The different symbols represent the populations as deduced from the cluster analysis. The black triangle marks the cluster center.}\label{Fig:phot cumul}
\end{center}
\end{figure*}
\subsection{Radial distribution}
Finally, we also performed the same cluster analysis on the full $uvby$ photometric catalogue after selecting all stars with $V$ magnitude brighter than 16.8 and located within a distance of 0.1\,mag from the chosen fiducials in all four colours. These criteria select 1100 stars, as shown in Fig.~\ref{Fig:phot selection}. The cumulative radial distribution for each population is given in the left panel of Fig.~\ref{Fig:phot cumul}, while the right panel displays the spatial distribution on the sky, with different symbols corresponding to the different populations.
A Kolmogorov-Smirnov test indicates a 4.5\% probability that the radial distributions of the primordial (green curve) and the extreme populations (blue curve) are drawn from the same parent distribution. The largest difference between the primordial and extreme population is found at a distance of roughly 3 arcmin from the cluster center. We also divided the full sample into subgroups: RGB stars ($16<V<13$), SGB stars ($16.8<V<16$), and proper motion members ($V > 13$; proper motion data from \citealt{Zloczewski2012}). We divided each group into five bins of radial distance to the cluster center. For each bin we calculated the mean distance to the cluster center for each population, and computed their number ratios. The radial distribution of the ratio between primordial and extreme stars is given in Fig.~\ref{Fig:rad dist}. Overplotted are the uncertainties given by the Poisson statistic, $\sigma = N_P/N_E\sqrt{1/N_P + 1/N_E}$, where $N_P$ and $N_E$ represent, respectively, the number of primordial and extreme stars within each bin. The dotted line indicates the half-mass radius $r_h = 2.34$\,arcmin \citep{Harris1996}.
We compare our radial distribution with that of \citet{Kravtsov2011}. They claim a strong radial segregation between the primordial and extreme populations, based on ground-based broadband $UBVI$ photometry of RGB stars. They conclude that RGB stars redder in ($U-B$) are centrally concentrated within the $r_h$, while the radial distribution of RGB stars bluer in ($U-B$) peaks outside this region. This rather extreme bimodality is incompatible with our results, which merely indicate that the primordial population falls off faster than the extreme population between a radial distance of 2 and 5 arcmin (see e.g. left panel of Fig.~\ref{Fig:phot cumul}, and Fig.~\ref{Fig:rad dist}) (within our sample). We do not see a radial segregation around the half-mass radius of the cluster. Our results are instead compatible with those of \citet{Milone2013}, who analysed multi-band {\sl Hubble Space Telescope} (HST) photometry, indicating no significant radial trend in the relative numbers of the three stellar populations within 6 arcmin of the cluster center. \citet{Milone2010} executed a Kolmogorov-Smirnov test on the radial distributions of different photometric populations on the SGB, indicating a 4\% probability that the primordial and extreme population have the same radial distribution, which they considered far from conclusive.\\
Recently, \citet{Kravtsov2014} finalised the study of the NGC\,6752 broadband photometry from \citet{Kravtsov2011}, including the radial distribution of blue horizontal branch (BHB) stars. They found that the brighter BHB (BBHB) stars are more centrally concentrated than the faint (FBHB), thus suggesting nitrogen enhanced red giants as the progenitors of BBHB stars, on the basis of qualitatively similar radial distributions. The ratio of BBHB to FBHB stars of \citet[Fig.~11]{Kravtsov2014} exhibits a similar morphology to our primordial-to-extreme star ratios in Fig.~\ref{Fig:rad dist}, with a minimum appearing at $R \approx 5$\,arcmin. They find also that the nitrogen enhanced extreme stars, RGBs redder in $U-B$ and the $U$-faint SGBs, are centrally concentrated (essentially within $r_h$) while their blue and $U$-faint counterparts exhibit flatter radial distributions in the field. This is contrary to our results indicating smooth distributions of extreme stars, while the primordial stars appear more centrally concentrated, falling off more rapidly outside $r_h$. \\
\begin{figure}
\begin{center}
\includegraphics[]{uvby_distribution_fullphot.eps}
\caption{Radial distributions of the ratio of primordial stars to extreme stars as function of the distance from the cluster center. The diamonds give the mean value of the radial bin whose range is indicated by the horizontal bar through the symbol. The dotted vertical line gives the half-mass radius of the cluster. }\label{Fig:rad dist}
\end{center}
\end{figure}
Both observed radial distributions can be argued for in terms of build-up scenarios for the cluster. On the one hand, 1D hydrodynamical simulations show that second-generation stars are formed more centrally than first-generation stars. This follows from the assumption that second-generation stars are made up of the ejecta from first-generation AGB stars. These ejecta accumulate in a cooling flow into the cluster core where the second-generation stars then form. Follow-up N-body simulations show that a large fraction of first-generation stars is lost early on. This because the cluster expands and early mass loss due to first-generation supernovae strips of the outer layers populated by first-generation stars \citep{Dercole2008,Vesperini2013}. During the dynamical evolutionary phase, both stellar generations mix and the ratio of second-to-first generation stars will settle into a morphology which is similar to the one observed by \citet{Kravtsov2014}. \\
On the other hand, we know that He-enhanced stars (i.e. second-generation stars) are slightly less massive than their He-normal counterparts (i.e. first-generation stars) when they evolve off the main sequence. The cluster will be driven toward equipartition of kinetic energy if one assumes that the dynamical evolutionary phase is dominated by two-body relaxation. As a result, the He-enhanced red giants can end up having a more extended distribution than He-normal ones after a Hubble time or several relaxation times. This is actually what we seem to find in our data sample: the primordial, i.e. He-normal, population falls off faster than their He-enhanced counterparts (see Fig.~\ref{Fig:rad dist}). There is no obvious explanation to reconcile both observations. We suggest further investigation of a consistent HST data set which extends to larger radial distance from the cluster center.
\section{Discussion}
We now return to the observed abundance trends and discuss their implications.
\begin{figure*}
\begin{center}
\includegraphics[]{diffusion.eps}
\caption{Averaged abundance trends from Figure~\ref{Fig:abundances} (green bullets), compared to the predictions from stellar structure models including atomic diffusion with additional mixing with two different efficiencies, at an age of 13.5\,Gyr. Horizontal, dashed lines represent the initial abundances of the models, which have been adjusted so that predictions match the observed abundance level of the coolest stars. The groups of three arrows in each panel indicate the influence on averaged abundances from increasing the effective temperature by 40\,K (left), surface gravity by 0.1\,dex (middle) and microturbulence by 0.1\,km\,s$^{-1}$\ (right) -- see also Table~\ref{Tab:abund sensitivity}.}\label{Fig:AD-trends}
\end{center}
\end{figure*}
\subsection{Diffusion trends}\label{subsect:AD}
Figure~\ref{Fig:AD-trends} compares the averaged trends of magnesium, calcium, titanium and iron from Fig.~\ref{Fig:abundances} (bullets) to predictions from stellar evolution models computed at the cluster's metallicity $\left[\mbox{Fe/H}\right] =-1.6$ (Paper~I). These models, introduced in Sect.\,\ref{sect:intro}, take into account the effects of atomic diffusion (AD), meaning radial transport of chemical elements, modelled from first principles, with the addition of a parametrised additional mixing (AddMix) of unknown physical origin.
AddMix is parametrised as a function of temperature and density (i.e., depth), and its efficiency is set by a free parameter representing a rescaling to the efficiency at a reference temperature $T_0$. Figure~\ref{Fig:AD-trends} compares observations to AD model predictions at two different efficiencies of AddMix, where the model T6.0 ($\log T_0 = 6.0$) represents a low efficiency of AddMix, while T6.2 ($\log T_0 = 6.2$) represents high efficiency.
Dashed horizontal lines indicate the initial model composition, which has been slightly adjusted so that predictions match the abundances of the cooler stars, where models coincide. This in turn is an effect of the convective envelope reaching deeper into the star, until it surpasses the region affected by AddMix.
Overall, we find good agreement between observations and predictions. The best correspondence is again found for the T6.2 model, as was the conclusion from the UVES analysis in Paper~I.
In Sect.~\ref{sec:clusteranalysis} we argued that the chemical populations in NGC\,6752 differ in helium fraction by $\Delta Y \sim0.03$. \citet{Lind2011a} showed, however, that such a change in helium fraction has negligible influence on $T_{\mbox{{\scriptsize eff}}}$\ and $\log g$\ at a given evolutionary phase. We thus assume that the variation in $Y$ between stars of different populations does not influence the predictions from our stellar-structure models (cf. the contrasting case of NGC\,6397, \citealt{Korn2007}). We also note that we find no significant abundance difference between stars of different populations in Ca, Ti, Cr, Fe and Ni, except for the evolutionary trends discussed here.
In the case of Mg, we find an insignificant tendency for the first-generation stars to exhibit a stronger trend than the second-generation stars, which is likely due to sampling of low-number statistics.
\\
The confirmation of the diffusion trends from Paper I brings further support to the assumption that AD is operational along the evolutionary sequence of NGC\,6752. Together with NGC\,6397 we now have two globular clusters in which effects of AD have been observed. Adding to this list the old, solar-metallicity open cluster M67, we can start to form a better picture about the r\^ole of AddMix and its behaviour with metallicity. Comparing the results in these three clusters with metallicities ranging from $\left[\mbox{Fe/H}\right] = -2.1$ (NGC\,6397) over $-1.6$ (NGC\,6752) to $+0.06$ (M67, \citealt{Onehag2014}), we find a decreasing effect of AD. One might have to make a differentiation between the metal-poor clusters and the solar-like ones as the degree of Addmix is found to increase with metallicity, i.e. decreasing the efficiency of AD, for the metal-poor clusters while AddMix seems unnecessary to explain the abundance trends in M67. As a final note we would like to comment on yet another metal-poor cluster which was studied by \citet{Mucciarelli2011}, namely M4 (NGC\,6121) at $\left[\mbox{Fe/H}\right] = -1.16$. Abundances of Fe and Li were derived for stars between the TOP and RGB. Although no trend in Fe was observed, the authors argue that the primordial lithium abundance (see Sect.~\ref{sec:evol-lithium}) can be reproduced assuming very efficient AddMix (T6.30). If this is true, the efficiency of AddMix is correlated with metallicity in the metal-poor regime. On the metal-rich side of the metallicity scale, however, AddMix does not seem to be required. One possible explanation is the fact that the outer convection zone extends further inside the star, surpassing the region where AddMix can operate. Continuing on this line of thought, one would expect the efficiency of AddMix to decrease with decreasing metallicity, as AddMix scales with the extent of the outer convection zone. Further analyses on globular clusters bridging the metallicity range between the metal-poor and the metal-rich regimes will shed light on the physical origin of AddMix. This is of utmost importance since any physical effects that alter the surface elemental composition of stars during their evolution must be taken into consideration in detailed studies of Galactic chemical evolution.
\subsection{Evolution of lithium} \label{sec:evol-lithium}
Figure~\ref{Fig:lithium} compares our derived lithium abundances to model predictions, with stars identified by their chemical population membership as determined in Sect.~\ref{sec:clusteranalysis}. Compared to the traditional Spite plateau identified among field stars, surface lithium abundances of these GC stars are known to be dramatically affected both by internal and external processes.
Firstly, surface layers of more evolved stars are diluted by essentially lithium-free processed material dredged up as the surface convection zone deepens along the subgiant branch. Secondly, star-to-star scatter is caused by processed material from short-lived massive stars mixing with the gas from which second-generation stars are forming, in what is known as intra-cluster pollution (see Sect.~\ref{sec:anticorr}).
Even selecting the least evolved, least polluted stars of our sample, the same type of mostly non-destructive surface depletion by gravitational settling as has been discussed for other elements is at work for lithium (as in all Spite plateau stars). As a matter of fact, lithium is the metal affected the most by gravitational settling.\\
Selecting the least processed stars in the cluster is however not straightforward. Surface dilution sets in after the TOP at $T_\text{eff} \approx 5900$\,K, which happens to coincide with the limit below which our photometric identification technique becomes less reliable (see Sect.~\ref{sec:clusteranalysis}). Only four TOP stars (out of 43) are identified as belonging to the primordial population. Correcting their measured lithium abundances for the effect of atomic diffusion predicted by the T6.2 model ($\sim +0.25$\,dex), we find an average initial lithium abundance $\abund{Li} = 2.52 \pm 0.03$ (statistical error).
\begin{figure}
\begin{center}
\includegraphics[]{li_cosm.eps}
\caption{Observed lithium abundances (coloured symbols as in Fig.~\ref{Fig:pop_abun}), compared to stellar evolution model predictions for two different efficiencies of AddMix (see text).
The initial abundance of the models (horizontal dashed line and shaded region), $\abund{Li} = 2.53 \pm 0.10$, compares well to the predicted primordial lithium abundance (dotted horizontal line and shaded region), $\abund{Li} = 2.69 \pm 0.04$.}\label{Fig:lithium}
\end{center}
\end{figure}
\begin{figure*}[ht]
\begin{center}
\includegraphics[]{initial_anticorr_li.eps}
\caption{Initial abundances (corrected for atomic diffusion and dredge-up) of lithium, compared to the anticorrelation elements magnesium (left), sodium (middle) and aluminium (right). The dotted line and shaded region refer to the predicted primordial abundance based on on BBN calibrated on the CMB (see text). The faintest ($V > 16.8$) stars have been excluded from the comparison. Different coloured symbols correspond to the chemical populations deduced from the cluster analysis.}\label{Fig:lithium_anticor}
\end{center}
\end{figure*}
Figure~\ref{Fig:pop_abun} indicates first-generation stars to be poor in sodium, with $\abund{Na} < 4.7$, in agreement with \citep{Carretta2013}. Selecting instead on this chemical criterion, matches 23 out of 30 TOP stars with observations of both sodium and lithium, indicating an initial abundance $\abund{Li} = 2.50 \pm 0.07$.
Applying these two selection criteria separately to the full sample of stars, including those affected by dredge-up, indicates $\abund{Li} = 2.54 \pm 0.05$ (25 stars) and $\abund{Li} = 2.52 \pm 0.06$ (52 stars). The strongest constraint is found when combining the two criteria, giving $\abund{Li} = 2.53 \pm 0.05$ (21 stars). We adopt this as our best estimate of the initial lithium abundance, but enlarge the error bar to 0.10\,dex in accordance with the uncertainty in selection and the systematic uncertainties discussed in Sect.~\ref{sec:lithium}.
These results are in good agreement with Paper~I, where an initial level of $\abund{Li} = 2.58\pm0.10$ was deduced from two TOP stars.\\
Continuing this line of reasoning, we compare in Figure~\ref{Fig:lithium_anticor} the initial abundances of lithium to magnesium, sodium and aluminium, after correcting for evolutionary effects (AD and dredge-up).
While magnesium abundances appear smoothly correlated with lithium, both sodium and aluminium exhibit a two-zone behaviour. In both Figure~\ref{Fig:lithium} and \ref{Fig:lithium_anticor}, the predicted primordial Li abundance based on Big Bang Nucleosynthesis (BBN) calibrated on the cosmological background radiation (CMB) as observed by Planck, is given by the shaded area at $\log\varepsilon$(Li)$_{\mbox{\scriptsize CMB+BBN}} = 2.69\pm0.04$ \citep{Coc2013}. The independent analysis of \citet{Steigman2013} prefers an insignificantly higher value of $2.72\pm0.04$.
\section{Conclusions}
Our differential 1D-LTE/NLTE based spectroscopic analysis of 194 stars observed at medium-high resolution, indicates weak ($\sim 0.1$\,dex) systematic abundance trends of heavy elements with evolutionary phase along the subgiant branch, in magnesium, calcium, titanium and iron. Although these trends are of low statistical significance taken individually, they are found to be in good agreement with those determined by independent methods in Paper~I, as well as predictions from stellar structure models including atomic diffusion with efficient additional mixing (the T6.2 model).
The trends are not likely to be caused by systematic errors on, e.g., effective temperature, as flattening the abundance trend in iron would require an increase of 200\,K on the turnoff point, implying an unrealistically large error in $\Delta\text{$T_{\mbox{{\scriptsize eff}}}$}(\text{TOP}-\text{bRGB})$ of 25\,\%.\\
Abundances of lithium, sodium, magnesium, aluminium and silicon exhibit intrinsic (anti)correlated variations, indicating the presence of multiple stellar populations in the cluster. It has previously been shown that these can be identified -- predicted -- using $uvby$ Str\"omgren photometry of stars on the RGB \citep{Carretta2011}. We have here extended this work to fainter, less evolved stars on the subgiant branch down to $V \approx 16.8$, ($T_{\rm eff} \approx 5900$\,K) close to the TOP. It seems there is a genuine colour index for every population with a given abundance pattern in the sense that the polluted chemical composition of second-generation stars makes them appear redder than first-generation stars. We have investigated the radial distribution of the different stellar populations by use of $uvby$ Str\"omgren photometry. We find a weakly significant indication of a weak radial segregation where the center of the globular cluster is slightly overpopulated by the primordial population, compared to the outskirts. \\
Combining photometric and spectroscopic information, we have attempted to identify the least polluted (primordial) stars in the cluster. Correcting their abundance patterns for the expected effects of atomic diffusion, we predict an initial lithium abundance of $\abund{Li} = 2.53 \pm 0.10$. This is in good agreement with the results of Paper~I, $\abund{Li} = 2.58 \pm 0.10$, and similar to the initial composition of the less massive globular cluster NGC\,6397, $\abund{Li} = 2.57 \pm 0.10$ \citep{Nordlander2012}. For both clusters, the initial lithium abundance compares reasonably well with Planck-calibrated predictions of standard BBN, $\log\varepsilon$(Li)$_{\mbox{\scriptsize CMB+BBN}} = 2.69\pm0.04$ \citep{Coc2013}. \\
Lithium is but one element affected by atomic diffusion and additional mixing, as all elements are affected to varying degree by these physical processes.
With recent constraints on atomic diffusion in M67 \citep{Onehag2014}, it is now important to bridge the gap between metal-poor globular clusters and the regime of solar-metallicity stars.
This will help shed light on the physical nature of additional mixing such that surface abundances of late-type stars can systematically be corrected for atomic diffusion and their initial abundances reliably recovered, for the benefit of such diverse science cases as studies of planet-hosting stars or the chemical evolution of the Galaxy.
\begin{acknowledgements}
We thank O. Richard for providing stellar-structure models, F.~Grundahl for providing the $uvby$ Str\"omgren photometric data, and Y.~Osorio for computing NLTE corrections for magnesium.
PG and AK thank the European Science Foundation for support in the framework of EuroGENESIS. TN and AK acknowledge support by the Swedish National Space Board.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Introduction}
\label{sec:intro}
Axions are hypothetical pseudoscalar particles that have been suggested as a
solution of the CP-violation problem in the strong interactions \cit
{P77,W78,Wl78}. Though axions arise as Nambu-Goldstone bosons and thus must
be fundamentally massless their interaction with gluons induces their mixing
with neutral pions. Axions thereby acquire a small mass which is
approximately given by \cite{BT78,KSS78,PBY87,GKR86}:
\begin{equation}
m_{a}=0.60\,\text{eV}\,\frac{10^{7}\text{GeV}}{f_{a}}, \label{ma}
\end{equation
where the unknoun constant $f_{a}$ with the dimension of an energy is the
axion decay constant. For a general review on axion physics see, e.g., \cit
{KSVZ,KSVZ1,DFSZ,DFSZ1,Kim,Cheng}. The axion phenomenology, in particular in relation
with the astrophysical processes, is largely discussed in \cit
{VZKC,R,R90,T90,R99,R08}.
Axions are a plausible candidate for the cold dark matter of the universe,
and a reasonable estimate of the axion mass (or, equivalently, the axion
decay constant) represents much interest. Over the years, various laboratory
experiments as well as astrophysical arguments have been used to constrain
the allowed range for $f_{a}$ or, equivalently, for the axion mass $m_{a}$.
Currently \cite{AS83,DF83}, cosmological arguments give $m_{a}>10^{-5}$ eV.
The most stringent upper limits on the axion mass derive from astrophysics.
Axions produced in hot astrophysical plasma can transport energy out of
stars. Strength of the axion coupling with normal matter and radiation is
bounded by the condition that stellar-evolution lifetimes or energy-loss
rates not conflict with observation. Such arguments are normally applied to
the physics of supernova explosions, where the dominant energy loss process
is the emission of neutrino pairs and axions in the nucleon bremsstrahlung
\cite{Br88,Bu88,R93,H00}. The limit from Supernova 1987A gives $m_{a}<0.01$
eV \cite{RS91,J96}. In works \cite{I84,U97} the thermal evolution of a
cooling neutron star was studied by including the axion emission in addition
to neutrino energy losses. The authors suggest the
upper limits on the axion mass of order $m_{a}<0.06-0.3$ eV by comparing the
theoretical curves with the ROSAT observational data for three pulsars: PSR
1055-52, Geminga and PSR 0656+14. Accuracy of such estimates substantially
depends on the assumptions of the matter equation of state and of the
effects of nucleon superfluidity which should be properly taken into
account. In the most cases the cooling scenario involves many parameters
which are poorly known.
The possibility of a more correct estimate has appeared following a
publication of analysed Chandra observations of the neutron star in
Cassiopeia A (Cas A NS) during 10 years \cite{H09,H10}. The authors found a
steady decline of the surface temperature, $T_{s}$, by about 4\% which they
interpret as a direct observation of Cas A NS cooling, the phenomenon which
has never been observed before for any isolated NS.
The decline is naturally explained if
neutrons have recently become superfluid (in $^{3}$P$_{2}$ triplet-state) in
the NS core, producing a splash of neutrino from pair breaking and formation
(PBF) processes\footnote{In Ref. \cite{St} the authors use the term Cooper
pair formation (CPF).} that currently accelerates the cooling \cite{St,P}.
The observed rapidity of the Cas A NS cooling implies that protons were
already in a superconducting $^{1}$S$_{0}$ singlet-state with a larger
critical temperature.
This scenario puts stringent constraints on poorly known properties of NS
cores. In particular, the density dependence of the temperature for the
onset of neutron superfluidity should have a wide peak with maximum
T_{c}(\rho )\approx (7-9)\times 10^{8}$~K.
\section{Neutrino and axion energy losses from superfluid NS core}
\label{sec:emis}
The neutrino pair emission caused by recombination of thermally broken Cooper
pairs \cite{FRS76,YKL} occurs through neutral weak currents generated by spin
fluctuations of the nucleons \cite{LP06,L10}.
Since the proton condensation occurs with a zeroth total spin
of a Cooper pair the spin fluctuations of the proton condensate are strongly
suppressed in the non-relativistic system \cite{FRS76}. As a result, the
dominating energy losses occur owing to the PBF neutrino radiation from
triplet pairing of neutrons, while the proton superfluidity quenches the
other neutrino reactions which efficiently operate in normal (nonsuperfluid)
nucleonic systems ($\bar{\nu}\nu $ bremsstrahlung, murca processes etc.)
Since the neutrino emission occurs mainly owing to neutron spin
fluctuations, the part of the interaction Hamiltonian relevant for PBF
processes is (we use natural units, $\hbar =c=k_{B}=1$)
\begin{equation}
\mathcal{H}_{\nu n}=-\frac{G_{F}C_{\mathsf{A}}}{2\sqrt{2}}\delta
_{\mu i}\left( \Psi ^{+}\hat{\sigma}_{i}\Psi \right)l^{\mu } , \label{Hnu}
\end{equation
where $l^{\mu }=\bar{\nu}\gamma ^{\mu }\left( 1-\gamma _{5}\right) \nu $ is
the neutrino current, $G_{F}=1.166\times 10^{-5}$ GeV$^{-2}$ is the Fermi
coupling constant, $\Psi $ is the nucleon field, $C_{\mathsf{A}}\simeq
1.26$ is the neutral-current axial-vector coupling constant of neutrons, and
$\hat{\sigma}_{i}$ are the Pauli spin matrices.
The dominant axion emission from a hot neutron star core is also caused by
spin fluctuations of non-relativistic neutrons. The corresponding Hamiltonian
density can be written in the form of derivative coupling
\begin{equation}
\mathcal{H}_{an}=\frac{c_{n}}{2f_{a}}\delta _{\mu i}\left( \Psi ^{+}\hat
\sigma}_{i}\Psi \right) \partial ^{\mu }a, \label{Ha}
\end{equation
where $c_{n}$ is the effective Peccei-Quinn charge of the neutron. This
dimensionless,
model-dependent coupling constant could range from $-0.05$ to $ 0.14$
\cite{M88,M89}.
The emission of neutrino pairs is kinematically possible owing to the
existence of a superfluid energy gap, which admits the quasiparticle
transitions with time-like momentum transfer $K=\left( \omega ,\mathbf{k
\right) $, as required by the final neutrino pair: $K=K_{1}+K_{2}$. \ The
energy-loss rate by $\bar{\nu}\nu $ emission caused by the neutron PBF
processes is given by the phase-space integral
\begin{equation}
Q_{\bar{\nu}\nu }\simeq \mathcal{N}_{\nu }\frac{G_{F}^{2}C_{\mathsf{A}}^{2}}
8}\int \;\frac{\omega }{1-\exp \frac{\omega }{T}}2\operatorname{Im}\Pi _
\mathsf{A}}^{\mu \nu }\left( \omega \right) \mathrm{Tr}\left( l_{\mu }l_{\nu
}^{\ast }\right) \frac{d^{3}k_{1}}{2\omega _{1}(2\pi )^{3}}\frac{d^{3}k_{2}}
2\omega _{2}(2\pi )^{3}}, \label{Q}
\end{equation
where $\mathcal{N}_{\nu }=3$ is the number of neutrino flavors, and
$\Pi _{\mathsf{A}}^{\mu \nu }$ is the retarded axial polarization tensor which
describes spin fluctuations in the neutron superfluid at temperature $T$.
The Fermi velocity is small in the nonrelativistic system, $V_{F}\ll 1$,
and we can study the neutrino energy losses in the lowest order over this small
parameter. Since the transferred space momentum comes in the polarization
functions in a combination $\mathbf{kV}_{F}\ll \omega ,\Delta $, one can
evaluate $\Pi _{\mathsf{A}}^{\mu \nu }$ in the limit $\mathbf{k}=0$.
After integration over the phase space of escaping neutrinos and
antineutrinos the total energy which is emitted into neutrino pairs per unit
volume and time is given by the following formula (See details, e.g., in
Ref. \cite{L01}):
\begin{equation}
Q_{\bar{\nu}\nu }=\frac{G_{F}^{2}C_{\mathsf{A}}^{2}}{64\pi ^{5}
\int_{0}^{\infty }d\omega \int\limits_{k<\omega }d^{3}q\frac{\omega }{1-\exp
\frac{\omega }{T}}\operatorname{Im}\Pi _{\mathsf{A}}^{\mu \nu }\left( \omega \right)
\left( K_{\mu }K_{\nu }-K^{2}g_{\mu \nu }\right) ~, \label{QQQ}
\end{equation
where we use a shortened notation $\Pi _{\mathsf{A}}^{\mu \nu }\left( \omega
\right) \equiv \Pi _{\mathsf{A}}^{\mu \nu }\left( \omega ,\mathbf{k
=0\right) $.
If now $K=\left(k,\mathbf{k}\right) $ denotes the axion four-momentum (we
ignore a small axion mass), the energy radiated per unit volume and time in
axions is given by the following phase-space integral
\begin{equation}
Q_{a}=\frac{1}{4}\frac{c_{n}^{2}}{f_{a}^{2}}\int \;\frac{k}{1-\exp \frac{k}{
}}2\operatorname{Im}\Pi _{\mathsf{A}}^{\mu \nu }\left( k\right) K_{\mu }K_{\nu
\frac{d^{3}k}{2k(2\pi )^{3}}. \label{rate1}
\end{equation
In the above, it was assumed that both axions and neutrinos can escape
freely from the medium so that final-state Pauli blocking factors can be
ignored.
The medium properties are embodied in a common function $\operatorname{Im}\Pi _
\mathsf{A}}^{\mu \nu }$ which is exactly the same for axion or neutrino
interactions because in Eqs. (\ref{QQQ}) and (\ref{rate1}) the global
coupling constants are explicitly pulled out. For the $^{3}$P$_{2}(m_{j}=0)$
pairing of neutrons this function is calculated in Ref. \cite{L10} with
taking into account of the ordinary and anomalous axial-vector vertices.
According to Eq. (93) of this work
\begin{align}
\operatorname{Im}\Pi _{\mathsf{A}}^{\mu \nu }\left( \omega \right) & =-\delta ^{\mu
i}\delta ^{\nu j}\frac{p_{F}M^{\ast }}{\pi ^{2}}\int d\mathbf{n}\left(
\delta _{ij}-\frac{\bar{b}_{i}\bar{b}_{j}}{\bar{b}^{2}}-\frac{3}{4}\left(
\delta _{ij}-\delta _{i3}\delta _{j3}\right) \right) \notag \\
& \times \Delta _{\mathbf{n}}^{2}\frac{\Theta \left( \omega ^{2}-4\Delta _
\mathbf{n}}^{2}\right) }{\omega \sqrt{\omega ^{2}-4\Delta _{\mathbf{n}}^{2}}
\tanh \frac{\omega }{4T}, \label{ImPi}
\end{align
where $p_{F}$ is the Fermi momentum of neutrons, $M^{\ast }\equiv
p_{F}/V_{F}$ is the neutron effective mass, and $\Theta \left( x\right) $ is the
Heaviside step-function.
For the $^{3}$P$_{2}(m_{j}=0)$ pairing the
normalised vector $\mathbf{\bar{b}}\left( \mathbf{n}\right) $ is defined as
\begin{equation}
\mathbf{\bar{b}}\left( \mathbf{n}\right) \equiv \sqrt{1/2}\left(
-n_{1},-n_{2},2n_{3}\right) . \label{bb}
\end{equation
Its angular dependence is represented by the unit vector $\mathbf{n=p}/p$
which defines the polar angles $\left( \theta ,\varphi \right) $ on the
Fermi surface
\begin{equation}
\mathbf{n=}\left( \sin \theta \cos \varphi ,\sin \theta \sin \varphi ,\cos
\theta \right) . \label{nn}
\end{equation}
The superfluid energy gap, generally defined by the relation
\begin{equation}
\Delta _{\mathbf{n}}^{2}=\mathbf{\bar{b}}^{2}\left( \mathbf{n}\right) \
\Delta ^{2}\left( \tau \right) , \label{Dns}
\end{equation
is anisotropic. It depends on the polar angle $\theta $ and on the relative
temperature $\tau \equiv T/T_{c}$. For the one component state $m_{j}=0$ one
has
\begin{equation}
\Delta _{\mathbf{n}}=\frac{1}{\sqrt{2}}\sqrt{1+3\cos ^{2}\theta }\ \Delta
\left( \tau \right) . \label{Dn}
\end{equation}
Insertion of Eq. (\ref{ImPi}) into Eqs. (\ref{QQQ}) and (\ref{rate1}) yields
the neutrino emissivity as given by Eq. (96) of Ref. \cite{L10}:
\begin{equation}
Q_{\nu }(m_{j}=0)\simeq \frac{2}{5\pi ^{5}}G_{F}^{2}C_{\mathsf{A
}^{2}p_{F}M^{\ast }T^{7}F_{4}\left( \frac{T}{T_{c}}\right) ~,
\label{Qnu}
\end{equation
and the axion emissivity
\begin{equation}
Q_{a}(m_{j}=0)=\frac{c_{n}^{2}}{f_{a}^{2}}\frac{2}{3\pi ^{3}}p_{F}M^{\ast
}T^{5}F_{2}\left( \frac{T}{T_{c}}\right) , \label{Qa}
\end{equation
wher
\begin{equation}
F_{l}\left( \tau \right) =\int \frac{d\mathbf{n}}{4\pi }\frac
\Delta _{\mathbf{n}}^{2}}{T^{2}}\int_{0}^{\infty }dx\frac{z^{l}}{\left( \exp
z+1\right) ^{2}} \label{J}
\end{equation
with $z=\sqrt{x^{2}+\Delta _{\mathbf{n}}^{2}/T^{2}}$. Details of the
numerical evaluation of this integral can be found in \cite{YKL,L10}.
\section{Mixed cooling by emission of axions and neutrino pairs}
\label{sec:mix}
Before proceeding to estimates of the axion radiation, let us note a few important
details of theoretical simulation of the CAS A NS neutrino cooling.
The authors of Ref. \cite{St} have reported that our Eq. (\ref{Qnu}) gives too slow
cooling. To achieve a better quantitative agreement of their simulation to the
observed data the neutrino energy losses were artifically enlarged in approximately
two times. This indicates that the thermal energy losses of Cas A NS are
approximately twice more intensive than neutrino losses given in Eq. (\ref{Qnu}).
Since currently there is no definitive explanation for this increase, we can speculate
that the additional energy losses from the superfluid core of the Cas A NS are
caused by axion emission, as described in Eq. (\ref{Qa}).
To get an idea of a compatibility of the axion emission with the CAS A NS observation
data let
us consider a simple model of cooling of the superfluid neutron core enclosed in a
thin envelope as typical for the NS. We assume that the bulk matter consists mostly
of $^{3}P_{2}$ superfluid neutrons with $m_{j}=0$.
In the temperature range which we are interested in, the thermal
luminosity of the surface is negligible in comparison to neutrino and
axion luminosities of the PBF processes in the NS core. In this case
the equation of global thermal balance \cite{gs80} reduces to
\begin{equation}
C(\widetilde{T}) \, { {\rm d} \widetilde{T} \over {\rm d} t}=
-L (\widetilde{T}).
\label{aa}
\end{equation}
Here $L (\widetilde{T})$ is the total
PBF luminosity of the star (redshifted to a distant observer), while
$C(\widetilde{T})$ is the stellar heat capacity. These quantities
are given by (see details in Refs. \cite{Y}):
\begin{eqnarray}
L (\widetilde{T})& = & \int {\rm d}V \, Q(T,\rho)
\exp(2 \Phi(r)),
\label{eq:Lnu} \\
C(\widetilde{T})& = & \int {\rm d}V \, C_V(T,\rho),
\label{C}
\end{eqnarray}
where $Q(T,\rho)$ is the total (neutrino + axion) emissivity,
$C_V (T,\rho)$ is the specific heat capacity,
${\rm d}V$ is the element of proper volume determined by
the appropriate metric function, and $\Phi(r)$ is the
metric function that determines gravitational redshift.
A thermally
relaxed star has an isothermal interior which extends from
the center to the heat blanketing envelope.
Taking into account the effects of General
Relativity (e.g., \cite{thorne77}), isothermality means
spatially constant redshifted internal temperature
\begin{equation}
\widetilde{T}(t)=T(r,t) \exp(\Phi(r)), \label{bb}
\end{equation}
while the local internal temperature $T(r,t)$,
depends on the radial coordinate $r$.
Given the strong dependence of the PBF processes on the temperature and the
density, the overall effect of simultaneous emission of neutrino pairs and axions can
only be assessed by complete calculations of the neutron star cooling which are
beyond the scope of this paper. A rough estimate can be made in a simplified model,
where the superfluid transition temperature $T_c$ is constant over the core.
In the temperature range of our interest, the specific heat is governed by
the neutron component (the contribution of electrons and strongly superfluid
protons is negligibly small) and can be described as
\begin{equation}
C\simeq\frac{1}{3}TR_{B}(T/T_c)
\int dVp_{F}M^{\ast }, \label{cc}
\end{equation}
where $R_B(T/T_c)$ is the superfluid reduction factor, as given in Eq. (18) of
Ref. \cite{YLS}.
Making use of Eqs. (\ref{Qnu}) and (\ref{Qa}) we obtain the PBF luminosity in
the form
\begin{equation}
L =\left[\frac{2}{5\pi ^{5}}G_{F}^{2}C_
\mathsf{A}}^{2}T^{7}F_{4}(T/T_c) +
\frac{c_{n}^{2}}{f_{a}^{2}}\frac{2}{3\pi ^{3}
T^{5}F_{2}(T/T_c)\right] \int dVp_{F}M^{\ast }e^{2\Phi ( r) }.
\label{dd}
\end{equation}
Insertion of Eqs. (\ref{bb}), (\ref{cc}) and (\ref{dd}) into Eq. (\ref{aa}) allows to obtain the
following equation for the non-redshifted temperature $T_{b}(t)\equiv T(r_b,t)$ at the edge
of the core or, equivalently, at the bottom of the envelope at $r=r_b$:
\begin{equation}
\frac{dT_{b}}{dt}=\frac{3\alpha }{R_{B}\left( T_{b}/T_{c}\right) }\left[ \frac{
}{5\pi ^{5}}G_{F}^{2}C_{\mathsf{A}}^{2}T_{b}^{6}F_{4}\left(
T_{b}/T_{c}\right) +\frac{c_{n}^{2}}{f_{a}^{2}}\frac{2}{3\pi ^{3}
T_{b}^{4}F_{2}\left( T_{b}/T_{c}\right) \right], \label{Tbeq}
\end{equation}
where the constant $\alpha\equiv\alpha (r_b)$ is defined as
\begin{equation}
\alpha \equiv \frac{\int dVp_{F}M^{\ast }e^{2\Phi \left( r\right) }}{\exp
\Phi \left( r_{b}\right) \int dVp_{F}M^{\ast }},
\label{alpha}
\end{equation}
and can be found from the CAS A NS observation data.
We convert the internal $T_b$ to the
observed effective surface temperature $T_{s}$ using (see, e.g., \cite{GPE,p06})
\begin{equation}
T_{s}/10^{6}\mathrm{K}\simeq (T_{b}/10^{8}\mathrm{K})^{0.55 }.
\label{Eq:TeTb}
\end{equation
This allows to compare the computed results with the observed (non-redshifted) CAS A NS
surface temperatures which are cataloged in Table 1 of Ref. \cite{St}.
\section{Results and discussion}
\label{sec:result}
For numerical estimate of the axion coupling strength to neutrons we designate
\begin{equation}
g=\frac{c_{n}^{2}}{f_{9}^{2}}, \label{g}
\end{equation}
with $f_{9}=f_{a}/\left( 10^{9}\text{GeV}\right)$, and consider $g$ as a free
parameter. Fig. \ref{fig:fig1} demonstrates the effect of mixed cooling of superfluid
neutron star with a constant $T_c$ over the core.
Two solid lines are the cooling curves for the simulated NS calculated at
$g=0.0$ and $g=0.16$.
\begin{figure}
\begin{center}
\includegraphics[width =1\textwidth]{fig1.eps}
\end{center}
\caption{(Color on line)
Cooling curves for a simulated CAS A NS consisting of a superfluid neutron core
and a low-mass blanketing envelope. $T_c$ is taken constant over the core.
Four curves correspond to the mixed (neutrino + axion) cooling at four values
$g=0$ ($T_c=7.55\times 10^8$~K), $g=0.1,\ 0.16$ and
$0.22$ ($T_c=7.2\times 10^8$~K). The points with error bars demonstrate
the observed surface temperatures cataloged in Table 1 of Ref. \cite{St}.
The inset shows the cooling curves but over larger range of ages. The lower
curve corresponds to the mixed cooling at $g=0.16$ while the upper curve
demonstrates cooling due to only neutrino emission artifically enhanced 2.1 times
as suggested in Ref. \cite{St}.}
\label{fig:fig1}
\end{figure}
The case $g=0$ describes the cooling caused by only the PBF neutrino emission
given in Eq. (\ref{Qnu}), with constant $T_c=7.55\times 10^8$ K. This curve
demonstrates too slow cooling and cannot explain the data. The case
$g=0.16$ agrees with the observations. This corresponds to the mixed
neutrino + axion radiation, as described by Eqs. (\ref{Qnu}) and (\ref{Qa}),
with $T_c=7.2\times 10^8$ K. The two dashed curves calculated at $g=0.1$ and
$g=0.22$ demonstrate that even a relatively small deviation off the value
$g=0.16$ results in substantial modification of the temperature profile and does
not allow to reproduce the observed cooling rate of the Cas A NS.
Thus we obtain $g\simeq 0.16$ or, equivalently,
\begin{equation}
\frac{c_{n}^{2}}{f_{a}^{2}}\simeq 1.6\times 10^{-19} \text{GeV}^{-2}.
\label{fa}
\end{equation}.
The same estimate immedeatly follows from a simple comparison of Eqs.
(\ref{Qnu}) and (\ref{Qa}) if one assumes that the axionic emissivity is
approximately equal to the neutrino emissivity. This gives
\begin{equation}
\frac{c_{n}^{2}}{f_{a}^{2}}\sim \frac{3}{5\pi ^{2}}G_{F}^{2}C_{\mathsf{A
}^{2}T^{2}\frac{F_{4}\left( \frac{T}{T_{c}}\right) }{F_{2}\left( \frac{
}{T_{c}}\right) }. \label{R1}
\end{equation
Inserting the typical values, $T_{c}\simeq 7.2\times 10^{8}\mathsf{K}$ and
$T\simeq 3.8\times 10^{8}\mathsf{K}$, we find $\tau \equiv T/T_{c}\simeq 0.53$ and
$F_{4}\left( \tau\right) /F_{2}\left( \tau \right) \simeq 10.4$.
Insertion of the above parameters into Eq (\ref{R1}) results in the estimate
given in Eq. (\ref{fa}).
One can use Eq. (\ref{ma}) to convert the decay constant $f_{a}$ to the axion mass
$m_{a}$. This yields
\begin{equation}
c_{n}^{2}m_{a}^{2}\sim 5.7\times 10^{-6}\,\text{eV}^2. \label{mc}
\end{equation
Unfortunately, the coupling constant $c_{n}$ depends on the axion model.
Given the QCD uncertainties of the hadronic axion models \cite{K85,S85,G82}, the
dimensionless constant $c_{n}$ could range from $-0.05$ to $ 0.14$.
While the canonical value $c_{n}=0.044$ is often used as generic examples, in
general $c_{n}$ is not known so that for fixed $ c_{n}^{2}m_{a}^{2}$ a broad
range of $m_a$ values is possible.
One should keep in mind that a strong cancelation of $c_{n}$ below
$c_{n}=0.044$ is also allowed.
In case of $c_{n}\rightarrow 0$ a powerfull PBF emission of axions is impossible.
This would mean that our assumption of the mixed cooling is invalid, and the PBF
neutrino losses are indeed at least two times larger than is predicted in Eq.
(\ref{Qnu}). Then the axion energy losses produce no noticeable modification of the
temperature profile of the CAS A NS, and one has to replace the Eq. (\ref{mc}) by
the inequality
\begin{equation}
c_{n}^{2}m_{a}^{2}\ll 5.7\times 10^{-6}\,\text{eV}^2. \label{mcl}
\end{equation
Can we discriminate the two cases from observations of the NS surface temperature?
As demonstrated in the insert in Fig. 1, the difference between the corresponding
theoretical cooling curves becomes discernable only after about 1000 years of cooling.
Perhaps future observation of the surface temperature of old neutron stars will help
to clarify the cooling mechanism.
Finally let us notice that our estimate of interaction of the hadronic axions with neutrons
has no analogies for a comparison. Previous astrophysical constraints was derived basically
for axions interacting simultaneously with neutrons and protons. In our case the proton
contribution is turned off due to large superfluid energy gap.
|
\section{Introduction}
Neutron stars are observed as several classes of self-gravitating systems: as radio and X-ray pulsars, as X-ray
bursters, as compact thermal X-ray sources in supernova remnants,
as rotating radio transients. In general, the structure of neutron stars and the
relation between the mass and the radius are determined by equations of
state (EoS) of dense matter.
The maximal mass of neutron star is still an open question. Recent
observations allows to estimate this limit at least as
$2M_{\odot}$: the well-measured limit of the pulsar PSR J1614-2230 is
$1.97M_{\odot}$ \cite{Demorest}, while, for the pulsar J0348+0432, it is
$2.01M_{\odot}$ \cite{Antoniadis}. Other examples of massive
neutron stars are Vela X-1 ($\sim 1.8M_{\odot}$ \cite{Rawls}) and
4U 1822-371 ($\sim 2M_{\odot}$, \cite{Munoz}). There are some
indications in favor of the existence of more massive neutron stars
with masses $\sim 2.4M_{\odot}$ (the possible masses of B1957+20
\cite{Kerk} and 4U 1700-377 \cite{Clark}) or even $\sim
2.7M_{\odot}$ (J1748-2021B \cite{Freire}).
It is interesting to note that for various EoS including hyperons,
the maximal mass limit for non-magnetic neutron stars is
considerably below than of the two-solar masses limit. The hyperonization process softens
EoS and then the maximal allowable mass results reduced
\cite{Glendenning,Glendenning-2,Schaffner,Vidana,Schulze}.
There are several ways to approach the solution of this problem (the so called ``{\it hyperon puzzle}'').
Firstly, the extensions of the simple model of hyperonic matter
(with three exchange meson fields - the so called
'$\rho\omega\sigma$'-model) allow to achieve the increasing of
maximal mass. Various approaches are proposed following this track. For
instance, larger hyperon-vector couplings (in comparison with
quark counting rule) require stiffness of the EoS
\cite{Hofmann,Rikovska,Sedrakian,Miyatsu-2}. Similar effect occurs
in model with chiral quark-meson coupling \cite{Miyatsu}. The
quartic vector-meson terms in the Lagrangian \cite{Bednarek} or the
inclusion of an additional vector-meson mediating repulsive
interaction amongst hyperons \cite{Weissenborn} also lead to the
increasing of the maximal mass limit. Authors of Ref. \cite{Whittenbury}
proposed an EoS with maximum mass $\sim 2.1M_{\odot}$ using the
quark-meson coupling model, which naturally incorporates hyperons
without additional parameters. A model with in-medium hyperon
interactions is considered in \cite{Wei}.
Another source for increasing the maximal mass limit is the
existence of strong magnetic fields inside the star.
The existence of soft gamma-ray repeaters and anomalous X-ray
pulsars can be linked to neutron stars with very
strong magnetic fields of the order $10^{15}$ G on the surface.
In these cases, the maximum magnetic field in the central regions of neutron star can
exceed $10^{18}$ G, according to the scalar virial
theorem. Such magnetic fields affect considerably the EoS for dense
matter and result in increasing the maximal mass of neutron
stars.
Various models of dense nuclear matter in presence of
strong magnetic fields have been considered in literature. The simplest model with
interacting $npe\mu$ gas is investigated in \cite{Lattimer}.
Models with hyperons and quarks are considered in
\cite{Rabhi}-\cite{Lopes}. It has been demonstrated that the Landau
quantization leads to the softening of the EoS for matter but account
for contributions of magnetic field into pressure and density. This fact leads, on the other hand,
to the stiffening of EoS.
Therefore neutron stars are very peculiar objects for testing theories
of matter at high density regimes and in strong magnetic fields.
It is interesting to note that data about neutron stars (mainly
mass-radius ($M-R$) relation) can be used for investigating
possible deviations from General Relativity (GR).
The initial motivation for studying modified gravity came from the
discovered accelerated expansion of the universe
confirmed by numerous independent observations. These observations
include Hubble diagram for Ia type supernovae \cite{Perlmutter,
Riess1,Riess2}, cosmic microvawe background radiation (CMBR) data
\cite{Spergel}, surveys of gravitational weak lensing
\cite{Schmidt} and data on Lyman alpha forest absorption lines
\cite{McDonald}.
This acceleration takes place at relatively small distances
(``Hubble flow'') and requires (in GR)
non-standard cosmic fluid (dark energy) filling the universe with negative pressure but not
clustered in large scale structure. The
nature of dark energy is unclear. Although from an observational
viewpoint, the so called $\Lambda$CDM model (where dark
energy is considered as Einstein Cosmological Constant) is in
agreement with data coming from observations there are various
problems and shortcomings at theoretical level. One of this issues is the
``smallness'' of cosmological constant i.e. the difference of 120
orders of magnitude between its observed value and the one
predicted by quantum field theory \cite{Weinberg}.
An alternative approach to dark energy problem consists of
extending of GR. In this case, the accelerated
expansion can be obtained without using ``dark energy'' but
enlarging the gravitational sector \cite{Capozziello1,
Capozziello2, Odintsov1, Turner, Odintsov-3,
Capozziello3,Capozziello_book, Capozziello4, Cruz}. Therefore
theories of modified gravity can be considered as real alternative
to GR.
The study of relativistic stars in modified gravity is interesting
from several reasons and could constitute a formidable probe for
such theories. Firstly one can reject some models that do not
allow the existence of stable star configurations \cite{Briscese,
Abdalla, Bamba, Kobayashi-Maeda, Nojiri5,Lang} (however one has to
note that stability can be achieved due to ``chameleon mechanism''
\cite{Tsujikawa, Upadhye-Hu} and may depend on the choice of the
EoS). Secondly there is the possibility for the existence of
new stellar structures, in the framework of modified gravity, escaping the standard stellar models.
The observation of such self-gravitating anomalous structures could provide strong
evidence for the Extended Gravity (see e.g. \cite{Laurentis, Laurentis2, Farinelli}).
The present paper is devoted to neutron stars with strong magnetic fields
in framework of analytic $f(R)$ gravity. Assuming a
simple model for strong interactions, one can obtain the EoS for
dense matter in magnetic field. Landau quantization, due to
magnetic fields, results to have significant effects. We consider the
cases of slowly and fast varying fields.
The paper is organized as follows. In Sec. II, we briefly
consider the the field equations for $f(R)$ gravity and the
modified Tolman--Oppenheimer--Volkoff (TOV) equations. Then
relativistic mean field theory for dense matter in strong magnetic
fields is presented (Sec.III).
In Sec. IV, the neutron star models for strong magnetic fields in
quadratic ($f(R)=R+\alpha R^2$) and cubic $f(R)=R+\alpha
R^3$ gravity are presented. The $M-R$ relation is derived and
compared with the one in GR. Conclusions and
outlooks are reported in Sec.V.
\section{Modified TOV equations in $f(R)$ gravity}
The action of $f(R)$-gravity is
\begin{equation}\label{action}
S=\frac{c^4}{16\pi G}\int d^4x \sqrt{-g}f(R) + S_{{\rm
matter}}\quad\,.
\end{equation}
It can be expressed as $f(R)=R+\alpha h(R)$.
The field equations are
\begin{equation}\label{field}
(1+\alpha h_{R})G_{\mu \nu }-\frac{1}{2}\alpha(h-h_{R}R)g_{\mu \nu
}-\alpha (\nabla _{\mu }\nabla _{\nu }-g_{\mu \nu }\Box
)h_{R}=8\pi G T_{\mu \nu }/c^{4}.
\end{equation}
Here $g$ is the determinant of the metric $g_{\mu\nu}$ and $S_{\rm
matter}$ is the action of the standard perfect fluid matter. The
Einstein tensor is $G_{\mu\nu}=R_{\mu\nu}-\frac12Rg_{\mu\nu}$ and
${\displaystyle h_R=\frac{dh}{dR}}$.
For star configurations, one can assumes a spherically symmetric
metric with two independent functions of radial coordinate, that
is:
\begin{equation}\label{metric}
ds^2= -e^{2\phi}c^2 dt^2 +e^{2\lambda}dr^2 +r^2 (d\theta^2
+\sin^2\theta d\phi^2).
\end{equation}
For the exterior solution, we assume a Schwarzschild metric.
Therefore it is convenient to define the variable \cite{Stephani,Cooney}
\begin{equation}\label{mass}
e^{-2\lambda}=1-\frac{2G M}{c^2 r}.
\end{equation}
The value of variable $M$ on the star surface is the gravitational
mass. For a perfect fluid, the energy-momentum tensor is
$T_{\mu\nu}=\mbox{diag}(e^{2\phi}\rho c^{2}, e^{2\lambda}P, r^2P,
r^{2}\sin^{2}\theta P)$, where $\rho$ is the matter density and $P$ is
the pressure. The field equations of interest are
\begin{eqnarray}
-8\pi G \rho/c^2 &=& -r^{-2} +e^{-2\lambda}(1-2r\lambda')r^{-2}
+\alpha h_R(-r^{-2} +e^{-2\lambda}(1-2r\lambda')r^{-2}) \nonumber \\
&& -\frac12\alpha(h-h_{R}R) +e^{-2\lambda}\alpha[h_R'r^{-1}(2-r\lambda')+h_R''] \label{f-tt},\\
8\pi G P/c^4 &=& -r^{-2} +e^{-2\lambda}(1+2r\phi')r^{-2}
+\alpha h_R(-r^{-2} +e^{-2\lambda}(1+2r\phi')r^{-2}) \nonumber \\
&& -\frac12\alpha(h-h_{R}R) +e^{-2\lambda}\alpha h_R'r^{-1}(2+r\phi'), \label{f-rr}
\end{eqnarray}
where $'\equiv d/dr$. The second TOV equation follows from the conservation law
$T_{\nu;\mu}^{\mu}=0$ and Eq.(\ref{f-rr}).
As result, the modified TOV equations can be written as
\cite{Astashenok}
\begin{equation}\label{TOV-1}
\left(1+\alpha h_{{R}}+\frac{1}{2}\alpha h'_{{R}}
r\right)\frac{dm}{dr}=4\pi{\rho}r^{2}-\frac{1}{4}\alpha r^2
\left[h-h_{{R}}{R}-2\left(1-\frac{2m}{r}\right)\left(\frac{2h'_{{R}}}{r}+h''_{{R}}\right)\right],
\end{equation}
\begin{equation}\label{TOV-2}
8\pi p=-2\left(1+\alpha
h_{{R}}\right)\frac{m}{r^{3}}-\left(1-\frac{2m}{r}\right)\left(\frac{2}{r}(1+\alpha
h_{{R}})+\alpha r_{g}^{2}
h'_{{R}}\right)({\rho}+p)^{-1}\frac{dp}{dr}-
\end{equation}
$$
-\frac{1}{2}\alpha
\left[h-h_{{R}}{R}-4\left(1-\frac{2m}{r}\right)\frac{h'_{{R}}}{r}\right],
$$
Here we have introduced the dimensionless variables $M=m M_{\odot},\quad
r\rightarrow r_{g}r, \quad \rho\rightarrow\rho
M_{\odot}/r_{g}^{3},\quad P\rightarrow p M_{\odot}c^{2}/r_{g}^{3},
\quad R\rightarrow {R}/r_{g}^{2}$, $\alpha
r_{g}^{2}h(R)\rightarrow \alpha h(R)$, where
$r_{g}=GM_{\odot}/c^{2}=1.47473$ km.
The third independent equation for Ricci curvature scalar is
\begin{equation}\label{TOV-3}
3\alpha
r_{g}^{2}\left[\left(\frac{2}{r}-\frac{3m}{r^{2}}-\frac{dm}{rdr}-\left(1-\frac{2m}{r}\right)\frac{dp}{(\rho+p)dr}\right)\frac{d}{dr}+
\left(1-\frac{2m}{r}\right)\frac{d^{2}}{dr^{2}}\right]h_{{R}}+\alpha
r_{g}^{2} h_{{R}}{R}-2\alpha r_{g}^{2} h-{R}=-8\pi({\rho}-3p)\,.
\end{equation}
Eqs. (\ref{TOV-1}), (\ref{TOV-2}), and (\ref{TOV-3}) can be solved
numerically for given EOS. In order to get solution, one can use perturbative
approach (see for details
\cite{Arapoglu,Alavirad,Astashenok,Astashenok-2}). In the framework of
perturbative approach, terms containing $h(R)$ are assumed to
be of first order in the small parameter $\alpha$, so all such
terms should be evaluated at ${\mathcal O}(\alpha)$ order. The
Ricci curvature scalar at zero order is $R^{(0)}=8\pi
(\rho^{(0)}-3p^{(0)})$. Therefore the deviation from GR strongly depends from the assumed form of EoS.
\section{Relativistic mean field theory for dense matter in presence of strong magnetic field}
Let us assume a simple model for describing nuclear matter in magnetic
field. The magnetic field $B$ is assumed along $z$-axis i.e. the
4-potential is $A^{mu}=(0,0,Bx,0)$. For nuclear matter consisting
of baryon octet ($b=$$p$, $n$, $\Lambda$, $\Sigma^{0,\pm}$,
$\Xi^{0,-}$) interacting with magnetic field and scalar $\sigma$,
isoscalar-vector $\omega_\mu$ and isovector-vector $\rho_\mu$ meson
fields and leptons ($l=$$e^{-}$, $\mu^{-}$), it is \cite{Typel}
\begin{equation}
\mathcal{L}=\sum_{b}\bar{\psi}_{b}\left(\gamma_{\mu}(i\partial^{\mu}-q_{b}A^{\mu}-g_{\omega
b}\omega^{\mu}-\frac{1}{2}g_{\rho
b}{\tau}\cdot{\rho}^{\mu})-(m_{b}-g_{\sigma
b}\sigma)\right)\psi_{b}+\sum_{l}\bar{\psi}_{l}\left(\gamma_{\mu}(i\partial^{\mu}-q_{l}A^{\mu})-m_{l}\right)\psi_{l}+
\end{equation}
$$
+\frac{1}{2}\left((\partial_{\mu}\sigma)^{2}-m^{2}_{\sigma}\sigma^{2}\right)-V(\sigma)-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}+\frac{1}{2}m^{2}_{\omega}\omega^{2}-\frac{1}{4}\omega_{\mu\nu}\omega^{\mu\nu}-
\frac{1}{4}{\rho}_{\mu\nu}{\rho}^{\mu\nu}+\frac{1}{2}m^{2}_{\rho}{\rho}_{\mu}^{2}.
$$
Here the mesonic and electromagnetic field strength tensors are
defined by the usual relations
$\omega_{\mu\nu}=\partial_{\mu}\omega_{\nu}-\partial_{\nu}\omega_{\mu}$,
$\rho_{\mu\nu}=\partial_{\mu}\rho_{\nu}-\partial_{\nu}\rho_{\mu}$,
$F_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}$. For
the sake of simplicity, we consider frozen-field configurations of
electromagnetic field. Also we neglect the anomalous magnetic
moments (AMM) of baryons and leptons because their effect is very
small. The strong interaction couplings $g_{b\sigma}$,
$g_{b\omega}$ and $g_{b\rho}$ depend from density. We use the
parameterization adopted in \cite{Typel}:
\begin{equation}
g_{i}(\rho)=g_{i0}f_{i}(x), x=\rho/\rho_{0}\,,
\end{equation}
where
$$f_{i}(x)=a_{i}\frac{1+b_{i}(x+d_{i})^2}{1+c_{i}(x+d_{i})^{2}}\,.$$
For the isovector field, it is
$$
g_{b\rho}=g_{b0}\exp[-a_{\rho}(x-1)].
$$
The values of constants $a_{i}$, $b_{i}$, $c_{i}$, $d_{i}$ are
given in \cite{Typel}. Using the mean-field approximation, one can obtain the following
equations for meson fields:
\begin{equation}\label{0}
m^{2}_{\sigma}\sigma+\frac{dV}{d\sigma}=\sum_{b}g_{\sigma b}
n_{b}^{s},\quad m^{2}_{\omega}\omega_{0}=\sum_{b}g_{\omega b}
n_{b},\quad m^{2}_{\omega}\rho_{03}=\sum_{b}g_{\rho b} n_{b}.
\end{equation}
Here $\sigma$, $\omega^{0}$, $\rho^{0}$ are the expectation values
of the meson fields in uniform matter. The quantities $n^{s}_{b}$,
$n_{b}$ are the scalar and vector baryon number densities,
correspondingly. The simplest scalar field potential is defined as
\begin{equation}
V(\sigma)=\frac{1}{3}b m_{N}(g_{\sigma
N}\sigma)^{3}+\frac{1}{4}c(g_{\sigma N}\sigma)^{4},
\end{equation}
where $b$ and $c$ are dimensionless constants. The values of
nucleon-meson couplings and parameters $b$, $c$ are given in Table
I. From the Dirac equations for charged and neutral baryons and
leptons, we have the energy spectra:
\begin{equation}
E^{b}_{\nu}=(k_{z}^{2}+m_{b}^{*2}+2\nu|q_{b}|B)^{1/2}+g_{\omega
b}\omega^{0}+\tau_{3b}g_{\rho b}\rho^{0}+\Sigma^{R}_{0},
\end{equation}
\begin{equation}
E^{b}=(k^{2}+m_{b}^{2})^{1/2}+g_{\omega
b}\omega^{0}+\tau_{3b}g_{\rho b}\rho^{0}+\Sigma^{R}_{0},
\end{equation}
\begin{equation}
E^{l}_{\nu}=(k_{z}^{2}+m_{l}^{2}+2\nu|q_{l}|B)^{1/2}.
\end{equation}
The number $\nu=n+1/2-sgn(q)s/2$ denotes the Landau levels of the
fermions with electric charge $q$, spin number $s=\pm 1$ for spin
up and spin down cases correspondingly. The spin degeneracy is
$g_{\nu}=1$ for lowest Landau level ($\nu=0$) and 2 for all other
levels. The effective mass for baryons is $m_{b}^{*}=m_{b}-g_{\sigma
b}\sigma$.
The rearrangement self-energy term is defined by
\begin{equation}
\Sigma^{R}_{0}=-\frac{\partial \ln g_{\sigma N}}{\partial
n}m^{2}_{\sigma}\sigma^{2}+\frac{\partial \ln g_{\omega
N}}{\partial n}m^{2}_{\omega}\omega^{2}_{0}+\frac{\partial \ln
g_{\rho N}}{\partial n}m^{2}_{\rho}\rho^{2}_{0}.
\end{equation}
Here $n=\sum_{b} n_{b}$.
The scalar densities for neutral baryons are \cite{Lattimer}
\begin{equation}
n^{s}_{b}=\frac{m_{b}^{*2}}{2\pi^2}\left(E^{b}_{f}k^{b}_{f}-
m_{b}^{*2}\ln\left|\frac{k^{b}_{f}+E^{b}_{f}}{m_{b}^{*}}\right|\right)
\end{equation}
and for charged baryons, it is
\begin{equation}
n^{s}_{b}=\frac{|q_{b}|B m_{b}^{*}}{2\pi^2}\sum_{\nu}
g_{\nu}\ln\left|\frac{k^{b}_{f,\nu}+E^{b}_{f}}{\sqrt{m_{b}^{*2}}+2\nu|q_{b}|B}\right|.
\end{equation}
For the vector densities for neutral baryons, we have
\begin{equation}
n_{b}=\frac{1}{3\pi^2} k^{b 3}_{f}.
\end{equation}
and for charged baryons and leptons, it is
\begin{equation}
n_{b,l}=\frac{|q_{b,l}|B}{2\pi^2} \sum_{\nu} g_{\nu}
k^{b,l}_{f,\nu}.
\end{equation}
Here $E_{f}^{b,l}$ is the Fermi energy. For charged baryon,
$E_{f}^{b}$ is related to the Fermi momentum $k_{f,\nu}^{b}$ as
$E_{f}^{b}=(k_{f}^{2}+m^{*2}_{b}+2\nu |q_{b}| B)^{1/2}$. For
neutral baryon, it is $E_{f}^{b}=(k_{f}^{2}+m^{*2}_{b})^{1/2}$. The
summation over $\nu$ terminates at value $\nu_{max}$ where the
square of Fermi momenta is still positive. For large magnetic
fields $B\sim 10^{18}$ G, only few Landau levels are occupied.
For hyperon-meson couplings there are no well-defined rule. One
can use for these constants quark counting rule \cite{Dover,Schafner}:
\begin{equation}
g_{\omega \Lambda}=g_{\omega \Sigma}=\frac{2}{3}g_{\omega N},
\quad g_{\omega \Xi}=\frac{1}{2}g_{\omega N},
\end{equation}
and
\begin{equation}
g_{\rho \Sigma}=2g_{\rho N}, \quad g_{\rho\Xi}=g_{\rho N}.
\end{equation}
Another choice is assuming that the fractions of
nucleon-meson couplings, i.e. $g_{iH}=x_{iH}g_{iN}$, is fixed. Here, it is
$x_{\sigma H}=x_{\rho H}=0.600$, $x_{\omega H}=0.653$, $x_{\rho
H}=0.6$ (see \cite{Rabhi}). We use this definition for further
calculations.
\begin{table}
\label{Table1}
\begin{centering}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|}
\hline
& $n_{s}$, & $-B/A$, & $M^{*}/M$ & $g_{0\sigma N}/m_{\sigma}$, & $g_{0\omega N}/m_{\omega}$, & $g_{0\rho N}/m_{\rho}$, & & \\
Model & fm$^{-3}$ & MeV & & fm & fm & fm & b & c \\
\hline
TW & 0.153 & 16.30 & 0.56 & 3.84901 & 3.34919 & 1.89354 & 0 & 0 \\
GM1 & 0.153 & 16.30 & 0.70 & 3.434 & 2.674 & 2.100 & 0.002947 & $-0.001070$ \\
GM2 & 0.153 & 16.30 & 0.78 & 3.025 & 2.195 & 2.189 & 0.003487 & $0.01328$ \\
GM3 & 0.153 & 16.30 & 0.78 & 3.151 & 2.195 & 2.189 & 0.008659 & $-0.002421$ \\
\hline
\end{tabular}
\caption{The nucleon-meson couplings and parameters of scalar
field potential for some models (GM1-3 - \cite{Glendenning}, TW -
\cite{Typel}). The nuclear saturation density $n_{s}$, the Dirac
effective mass $M^{*}$ and the binding energy ($B/A$) are also
given.}
\end{centering}
\end{table}
For chemical potential of baryons and leptons, one has
$$
\mu_{b}=E^{f}_{b}+g_{\omega b} \omega_{0}+g_{\rho
b}I_{3}\rho_{03}+\Sigma_{0}^{R},\quad \mu_{l}=E^{f}_{l}.
$$
The following conditions should be
imposed on the matter in order to obtain the EoS with the following properties:\\
\\ (i) baryon number
conservation:
\begin{equation}\label{1}
\sum_{b}n_{b}=n,
\end{equation}
(ii) charge neutrality:
\begin{equation}\label{2}
\sum_{i} q_{i}n_{i}=0,\quad i=b,l,
\end{equation}
(iii) beta-equilibrium conditions:
\begin{equation}\label{3}
\mu_{n}=\mu_{\Lambda}=\mu_{\Xi^{0}}=\mu_{\Sigma^{0}}, \quad
\mu_{p}=\mu_{\Sigma^{+}}=\mu_{n}-\mu_{e},\quad
\mu_{\Sigma^{-}}=\mu_{\Xi^{-}}=\mu_{n}+\mu_{e},\quad
\mu_{e}=\mu_{\mu}.
\end{equation}
At given $n$, Eqs. (\ref{0})-(\ref{3}) can be solved
numerically and one can find the Fermi energy for particles
and meson fields. Resulting matter energy density is
\begin{equation}
\epsilon_{m}=\sum_{b} \epsilon_{b}+\sum_{l}
\epsilon_{l}+\frac{1}{2}m_{\sigma}^{2}+\frac{1}{2}m_{\omega}^{2}+\frac{1}{2}m_{\rho}\rho^{2}_{0}+U(\sigma).
\end{equation}
The energy density for charged baryons is
\begin{equation}\label{encb}
\epsilon^{c}_{b}=\frac{|q_{b}|B}{4\pi^2}\sum_{\nu}g_{\nu}\left(k^{b}_{f,\nu}E^{b}_{f}+
(m_{b}^{*2}+2\nu|q_{b}|B)\ln\left|\frac{k^{b}_{f,\nu}+E^{b}_{f}}{\sqrt{m_{b}^{*2}+2\nu|q_{b}|B}}\right|\right)
\end{equation}
and for neutral baryons
$$
\epsilon^{n}_{b}=\frac{1}{4\pi^2}\left[k^{b}_{f}(E^{b}_{f})^{3}-\frac{1}{2}m_{b}^{*}\left(m_{b}^{*}k^{b}_{f}E^{b}_{f}
+m_{b}^{*3}\ln\left|\frac{k^{b}_{f}+E^{b}_{f}}{m_{b}^{*}}\right|\right)\right].
$$
The expression of energy density for leptons can be obtained from
(\ref{encb}) by changing $m_{b}^{*}\rightarrow m_{l}$. The pressure
of dense matter is defined as
$$
p=\sum_{b} p_{b}+\sum_{l}
p_{l}-\frac{1}{2}m_{\sigma}^{2}+\frac{1}{2}m_{\omega}^{2}+\frac{1}{2}m_{\rho}\rho^{2}_{0}-U(\sigma)+\Sigma^{R}_{0}.
$$
where pressure for charged baryons is
\begin{equation}
p^{c}_{n}=\frac{|q_{b}|B}{12\pi^2}\sum_{\nu}g_{\nu}\left(k^{b}_{f,\nu}E^{b}_{f}-
(m_{b}^{*2}+2\nu|q_{b}|B)\ln\left|\frac{k^{b}_{f,\nu}+E^{b}_{f}}{\sqrt{m_{b}^{*2}+2\nu|q_{b}|B}}\right|\right)
\end{equation}
and for neutral baryons
\begin{equation}
p^{n}_{b}=\frac{1}{12\pi^2}\left[k^{b}_{f}(E^{b}_{f})^{3}-\frac{3}{2}m_{b}^{*}\left(m_{b}^{*}k^{b}_{f}E^{b}_{f}
-m_{b}^{*3}\ln\left|\frac{k^{b}_{f}+E^{b}_{f}}{m_{b}^{*}}\right|\right)\right].
\end{equation}
In order to obtain the EoS, one needs to add the contribution of magnetic
field, that is
\begin{equation}
\epsilon=\epsilon_{m}+\frac{B^2}{8\pi},\quad
p=p_{m}+\frac{B^2}{8\pi}.
\end{equation}
We use a model where the magnetic field
depends from baryon density only. The parameterization proposed in \cite{Rabhi}, \cite{Ryu-2} has the form
\begin{equation}
B=B_{s}+B_{0}\left[1-\exp\left(-\beta(n/n_{s})^{\gamma}\right)\right],
\end{equation}
where $B_s$ is the magnetic field on the star surface ($10^{15}$ G).
For parameters $\gamma$ and $\beta$, one takes the values
$\gamma=2$, $\beta=0.05$ (slowly varying field) and $\gamma=3$,
$\beta=0.02$ (fast varying field). The value $B_{0}$ is convenient
to give in units of critical field for electron $B_{c}=4.414\times
10^{13}$ G.
All these considerations can be applied to models where curvature corrections appear in the TOV equations. Specifically, we adopt quadratic and cubic corrections.
\begin{table}
\label{Table2}
\begin{centering}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
$B_{0}$, & $\alpha$, & $M_{max}$, & $R$, & $E_{c}$, & $B_{c}$, \\
$10^5$ G & $10^{9}$ cm$^{2}$ & $M_{\odot}$ & km & GeV/fm$^{3}$ & $10^{18}$ G \\
\hline
& 0 & 1.51 & 10.00 & 1.61 & 0 \\
0 & $-5$ & 1.55 & 10.00 & 1.61 & 0 \\
& $5$ & 1.46 & 10.05 & 1.49 & 0 \\
\hline
& 0 & 2.21 & 11.69 & 1.17 & 3.38 \\
1 & $-5$ & 2.30 & 11.58 & 1.27 & 3.56 \\
& 5 & 2.14 & 11.82 & 1.09 & 3.20 \\
\hline
& 0 & 2.80 & 13.99 & 0.79 & 3.50 \\
2 & $-5$ & 2.91 & 13.44 & 0.97 & 3.93 \\
& $5$ & 2.71 & 14.31 & 0.68 & 3.06 \\
\hline
& 0 & 3.21 & 15.67 & 0.63 & 3.29 \\
3 & $-5$ & 3.33 & 15.24 & 0.68 & 3.75 \\
& 5 & 3.11 & 16.03 & 0.54 & 2.87 \\
\hline
\end{tabular}
\caption{Neutron star properties using TW model for quadratic
gravity (slowly varying field). The energy density ($E_{c}$) and
magnetic field ($B_{c}$) in the center for neutron star with
maximal mass are given.}
\end{centering}
\end{table}
\begin{table}
\label{Table3}
\begin{centering}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
$B_{0}$, & $\alpha$, & $M_{max}$, & $R$, & $E_{c}$, & $B_{c}$, \\
$10^5$ & $10^{9}$ cm$^{2}$ & $M_{\odot}$ & km & GeV/fm$^{3}$ & $10^{18}$ G \\
\hline
& 0 & 2.32 & 11.50 & 1.17 & 3.96 \\
1 & $-5$ & 2.44 & 11.47 & 1.22 & 4.36 \\
& 5 & 2.23 & 11.66 & 1.04 & 3.66 \\
\hline
& 0 & 2.73 & 12.44 & 0.93 & 4.20 \\
2 & $-5$ & 2.89 & 12.81 & 0.97 & 4.34 \\
& $5$ & 2.60 & 13.31 & 0.71 & 3.29 \\
\hline
& 0 & 2.98 & 13.78 & 0.76 & 3.87 \\
3 & $-5$ & 3.12 & 13.81 & 0.83 & 4.12 \\
& 5 & 2.84 & 14.06 & 0.68 & 3.50 \\
\hline
\end{tabular}
\caption{Neutron star properties using TW model for quadratic
gravity (fast varying field).}
\end{centering}
\end{table}
\begin{table}
\label{Table4}
\begin{centering}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
$B_{0}$, & $\alpha$, & $M_{max}$, & $R$, & $E_{c}$, & $B_{c}$, \\
$10^5$ & $10^{9}$ cm$^{2}$ & $M_{\odot}$ & km & GeV/fm$^{3}$ & $10^{18}$ G \\
\hline
2 & $10$ & 2.30 & 11.82 & 1.49 & 5.97 \\
\hline
3 & $10$ & 2.52 & 12.86 & 1.40 & 6.08 \\
\hline
\end{tabular}
\caption{Compact star properties on the second ``branch of
stability'' using TW model for quadratic gravity (fast varying
field). The magnetic field at the center can exceed $6\times
10^{18}$ G and the central energy density is approximately twice
than in GR.}
\end{centering}
\end{table}
\begin{center}
\begin{figure}
\includegraphics[scale=1.1]{MR0.eps}\\
\caption{The mass-radius diagram in model $f(R)=R+\alpha R^2$ for two values of $\alpha$ without magnetic field in comparison with GR.}
\end{figure}
\begin{figure}
\includegraphics[scale=1.1]{MRB.eps}\\
\caption{The mass-radius diagram in model $f(R)=R+\alpha R^2$ and in GR for slowly varying magnetic field ($B_{0}=1,\quad 2,\quad 3\times 10^5$).
The cases $\alpha=-5\times 10^9$, $0$, $5\times 10^9$ cm$^2$
correspond to dotted, thick and thin lines correspondingly.}
\end{figure}
\begin{figure}
\includegraphics[scale=1.1]{MRBF.eps}\\
\includegraphics[scale=1.1]{MRBF2.eps}\\
\caption{The Mass-Radius diagram in model $f(R)=R+\alpha R^2$ and in GR for fast varying field.
On upper panel the cases $\alpha=-5\times 10^9$, $0$, $5\times 10^9$ cm$^2$
correspond to dotted, thick and thin lines correspondingly. On lower panel the mass radius relation for $\alpha=10^{10}$ cm$^2$ is given (dotted lines).
The second ``branch of stability'' with more compact (in comparison with GR) neutron stars exists.}
\end{figure}
\end{center}
\begin{table}
\label{Table5}
\begin{centering}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
$B_{0}$, & $\beta$, & $M_{max}$, & $R$, & $E_{c}$, & $B_{c}$, \\
$10^5$ & $r_{g}^{4}$ & $M_{\odot}$ & km & GeV/fm$^{3}$ & $10^{18}$ G \\
\hline
& 0 & 1.51 & 10.00 & 1.61 & 0 \\
0 & $-50$ & 2.11 & 9.87 & 1.27 & 0 \\
& $-75$ & 2.45 & 10.02 & 1.22 & 0 \\
\hline
& 0 & 2.21 & 11.69 & 1.17 & 3.38 \\
1 & $-50$ & 2.70 & 11.07 & 1.67 & 4.09 \\
& $-75$ & 3.10 & 10.97 & 1.81 & 4.19 \\
\hline
& 0 & 2.80 & 13.99 & 0.79 & 3.50 \\
2 & $-25$ & 3.07 & 13.52 & 0.93 & 3.97 \\
& $-50$ & 3.29 & 13.78 & 0.83 & 3.62 \\
\hline
\end{tabular}
\caption{Neutron star properties using TW model for cubic gravity
for several values of $\beta$ (in units of $r_{g}^4=4.73\times
10^{21}$ cm$^4$) for slowly varying magnetic field.}
\end{centering}
\end{table}
\begin{table}
\label{Table6}
\begin{centering}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
$B_{0}$, & $\beta$, & $M_{max}/M_{\odot}$ & $R$, & $E_{c}$, & $B_{c}$, \\
$10^5$ & $r_{g}^{4}$ & & km & GeV/fm$^{3}$ & $10^{18}$ G \\
\hline
& 0 & 2.32 & 11.50 & 1.17 & 3.96 \\
& $-25$ & 2.73 & 11.10 & 1.27 & 4.14 \\
1 & $-50$ & 3.14 & 11.19 & 1.17 & 3.96 \\
& $-75$ & 3.64 & 11.18 & 1.17 & 3.96 \\
\hline
& 0 & 2.73 & 12.44 & 0.93 & 4.20 \\
2 & $-25$ & 3.24 & 12.87 & 0.86 & 3.93 \\
& $-50$ & 3.71 & 13.09 & 0.76 & 3.54 \\
\hline
\end{tabular}
\caption{Compact star properties using TW model for cubic gravity
for several values of $\beta$ for fast varying magnetic field.}
\end{centering}
\end{table}
\begin{figure}
\includegraphics[scale=1.1]{MR3s.eps}\\
\includegraphics[scale=1.1]{MR3f.eps}\\
\caption{The mass-radius diagram in model $f(R)=R+\beta R^3$ and in GR for slowly (upper panel) and fast (lower panel) varying field. One can see that
for the deviation of M-R relation from GR is smaller for larger values of
$B_{0}$.}
\end{figure}
\section{The cases of quadratic and cubic curvature corrections}
Let us firstly take into account models with quadratic gravity corrections, that is
\begin{equation}
f(R)=R+\alpha R^2.
\end{equation}
Neutron stars with strong magnetic field in quadratic gravity
was considered in \cite{EKSI} for relatively stiff EoS based on
model with five meson fields. We consider the quadratic gravity
case for EoS based on the above described model. For our calculations, the
Typel-Wolter (TW) parametrization is used.
Let us note the following feature: for $B=0$, the mass of neutron star
increases with decreasing $\alpha$ for the
various radii (see Fig. 1). For strong magnetic field, the $M-R$
relation for $M>0.7M_{\odot}$ differs from such in GR only for masses close to the maximal one (see Figs. 2,
3). Another interesting feature appears for fast varying field. At
high central densities, a second ``branch'' of stability can
exists (Fig. 3, lower panel).
It is interesting to note that similar effects take place for
non-magnetic neutron stars in the framework of a model like $f(R)=R+\alpha
R^2(1+\gamma R)$ \cite{Astashenok}. The stabilization of star
configurations occurs thanks to the cubic term.
The maximal masses and corresponding radii are given in Tables II,
III for some values of $\alpha$ and $B_{0}$. The maximal value of
central density (and therefore magnetic field) decreases with
increasing $\alpha$.
The parameters for compact (in comparison with GR) neutron stars
on second ``branch of stability'' are given in Table IV.
For modified gravity with cubic term, $f(R)=R+\beta R^{3}$, the
maximal value of neutron star mass for given EoS increases for
$\beta<0$ (Fig. 4). Some results are given in Tables V, VI. The
maximal mass of neutron star can exceed $3M_{\odot}$. One can note
that stars with magnetic field and cubic curvature corrections result stable for central energy density close to $\sim 1.8$
GeV/fm$^3$.
In principle, calculations show that, for EoS based on GM2-GM3 parameterizations,
we have similar results for models with
$f(R)=R+\alpha R^2$ and $f(R)=R+\beta R^3$. For more stiff EoS the
deviation from GR is larger.
\section{Conclusions and perspectives}
We presented neutron star models with strong
magnetic fields in the framework of power-law $f(R)$ gravity models. For
describing dense matter in magnetic field, a model with
baryon octet interacting through $\sigma$$\rho$$\omega$-fields is
used.
Although the softening of nucleon EoS, due to hyperonization,
leads to the decrease of the upper limit mass of neutron star, the
strong magnetic field can increase considerably the maximal mass
of star.
In particular, we investigated the effect of strong magnetic field in models of
quadratic, $f(R)=R+\alpha R^2$, and cubic, $f(R)=R+\alpha R^3$,
gravity. For large fields, the $M-R$ relation differs considerably
from such in GR only for stars with masses close
to maximal. Another interesting feature is the possible existence of
more compact stable stars with extremely large fields ($\sim
6\times 10^{18}$ G instead $\sim 4\times 10^{18}$ G in GR) in central regions of star. Due to the cubic term, the
significant increasing of maximal mass ($M_{max}>3M_{\odot}$) is
possible. The central energy density can exceed $\sim 1.8$
GeV/fm$^3$.
However, it is worth stressing that the $f(R)$ models considered here can be related to the presence of strong gravitational fields where higher order curvature terms can emerge. Their origin is related to the effective actions of quantum field theory formulated in curved spacetime \cite{buchbinder, birrel}. In the extreme field of neutron stars, it is realistic supposing the emergence of curvature corrections that improve the pressure effects and could explain supermassive self-gravitating systems.
As a next step, we will consider models of self-bounded quark
stars and hybrid stars with quark cores. The EoS for quark matter
(without magnetic field) is close to $p\sim \frac{1}{3} \rho c^2$
and therefore, in the framework of perturbative approach, the deviations
from GR occur only for very large values of
$\alpha$ in comparison with the above considered in quadratic
gravity. However, for large magnetic fields, considerable effects can
be induced on EoS and therefore the modified gravity effects can
appear.
\acknowledgments
This work is supported in part by projects 14-02-31100 (RFBR,
Russia) (AVA), by MINECO (Spain), FIS2010-15640 and by MES project
TSPU-139 (Russia) (SDO). SC is supported by INFN ({\it iniziative
specifiche} TEONGRAV and QGSKY).
|
\section{Introduction}
\label{intro}
The continuum X-ray and radio emission and mass of accreting black
holes are empirically correlated \citep{merlonietal03, fkm04}. The
correlation, frequently called the ``fundamental plane of black hole
accretion,'' relates mass accretion rates, probed by the X-ray
luminosity, and jet or outflow power, probed by the radio luminosity,
at a given mass of the black hole, which sets the size scale of the
accretion-disk--jet system. The relation was first seen in
stellar-mass black holes, which have a narrow range in mass
\citep{2003MNRAS.344...60G}, and then extended to all black holes
\citep{merlonietal03, fkm04}, covering over 8 orders of magnitude in
mass, 14 orders of magnitude in radio luminosity, and 12 orders of
magnitude in X-ray luminosity.
Despite the wide range that the fundamental plane covers, there is
room to question its universality. For example, using a larger sample
of stellar-mass black hole data, \citet{2012MNRAS.423..590G} found
that the radio--X-ray correlation was best explained by two tracks,
with one source that transitions from one track to the other.
Restricting to only black holes with dynamical mass measurements,
\citet{2009ApJ...706..404G} found different results when including and
excluding stellar-mass black holes ($L_R\sim M^{0.78}L_X^{0.67}$ and
$L_R\sim M^{2.08}L_X^{0.50}$, respectively).
There are two potential explanations for the separate relations found
by \citet{2009ApJ...706..404G}. The first is that supermassive black
holes (SMBHs) and stellar-mass black holes do occupy the same fundamental
plane, and the apparent difference is a
result of a small sample size and limited dynamic range. The second
is that SMBHs and supermassive black holes each
operate under different physical conditions and follow their own
relation, and fits to combined samples only appear to produce a single
relation because of the large logarithmic range in values. Either
empirical result would provide insight to the physics at play.
In this Letter, we report an observational study designed to test
which empirical relation is better. We selected a sample of active galactic nuclei (AGNs) with
masses in the range that is best suited to make an observational
distinction between the two relations (Section \ref{sample}). We
obtained new \emph{Chandra} and Karl G.\ Jansky Very Large Array (VLA)
observations and supplemented these with archival data to put them on the
fundamental plane (Sections \ref{xrayobs} and \ref{radioobs}). We
find that the single plane is a better predictor of the observed radio
and X-ray fluxes, and we discuss the implications of our results
(Section \ref{discussion}). Throughout this Letter we assume $H_0=70\ \units{km\ s^{-1}\ Mpc^{-1}}$, $\Omega_M=0.3$, and
$\Omega_\Lambda=0.7$.
\section{Observations}
\label{observations}
\subsection{Sample Selection}
\label{sample}
To test whether an SMBH-only relation or an all-black-holes relation
was better, we assembled a sample of 13 low-mass AGNs from the sample
of \citet{gh07}. Black hole masses ($\mbh$) come from the virial
(i.e., single-epoch) method using broad H$\alpha$ emission lines and
an assumed relation between the radius of the broad-line region and
continuum luminosity of the AGN, inferred from the broad H$\alpha$
luminosity. The parent sample was defined to have
$\mbh<10^{6.3}\ \msun$. Some of the parent sample were flagged with c
by \citet{gh07} to indicate that the broad-line emission may not have
been robustly detected. The c-sample sources may therefore have less
robust mass estimates. As only one c-sample source is detected in
both radio and X-ray, our fundamental-plane analysis and conclusions
below are not affected.
Out of 229 low-mass AGNs, we selected 13 for which the X-ray and radio
fluxes could be measured in a reasonable exposure as predicted by the
\citet{2009ApJ...706..404G} fundamental plane. To determine the
suitability of potential sources for our new X-ray and radio
observations (Sections \ref{xrayobs} and \ref{radioobs}), we used
information from existing optical observations. For X-ray
detectability, we assumed the bolometric luminosities from
\citet{gh07} and bolometric correction from
\citet{2007MNRAS.381.1235V}. For VLA detectability, we assumed the
more conservative of the two \citet{2009ApJ...706..404G} fundamental
plane predictions.
We included both sources that have VLA FIRST
\citep{1997ApJ...475..479W} 1.4 GHz detections and those that
do not, so as not to bias ourselves to those brightest in radio
wavelengths. As discussed below, VLA FIRST detection was not a good
predictor of 8.5 GHz continuum emission.
To reduce scatter related to source variability, we scheduled
\emph{Chandra} and VLA observations contemporaneously. We also used
archival X-ray data when possible. Table \ref{t:sample} lists basic
properties---including \citet{gh07} identification number, which we use here---of all low-mass AGNs in our sample. We
have updated the masses to use the most recent single-epoch H$\alpha$
scaling relations \citep{2005ApJ...630..122G, 2013ApJ...767..149B,
2013ApJ...775..116R} and to assume a scaling factor of
$\epsilon=4.3/4$ \citep{2013ApJ...773...90G}, which uses the approximation
$V_\mathrm{FWHM}=2\sigma$. This makes a typical difference of $<0.1$
dex compared to \citet{gh07}, much smaller
than the estimated 0.5 dex systematic uncertainty.
\kgtabbeg{r@{\extracolsep{0pt}}llrrrr}
\footnotesize
\tablecaption{Sample of Small AGN}
\tablewidth{\columnwidth}
\tablehead{
\multicolumn{2}{c}{GH ID} &
\colhead{SDSS} &
\colhead{$z$} &
\colhead{$D_L$} &
\colhead{$\log(M/\msun)$} &
\colhead{$\Delta T_\mathrm{obs}$}
}
\startdata
47& & J082443.28+295923.5 & 0.025 & 109 & 5.70 & 1890 \\
69& & J091449.05+085321.1 & 0.140 & 661 & 6.30 & 176 \\
87& & J101246.49+061604.7 & 0.078 & 353 & 6.22 & 60 \\
94&$^\mathrm{c}$ & J103234.85+650227.9 & 0.006 & 26 & 5.80 & \dots \\
101&$^\mathrm{c}$ & J105108.81+605957.2 & 0.082 & 373 & 6.27 & \dots \\
106& & J110501.97+594103.6 & 0.034 & 149 & 5.58 & 822 \\
119&$^\mathrm{c}$ & J112637.74+513423.0 & 0.026 & 114 & 6.16 & \dots \\
140&$^\mathrm{c}$ & J121629.13+601823.5 & 0.060 & 269 & 6.18 & 76 \\
146& & J124035.81$-$002919.4 & 0.081 & 368 & 6.35 & 2824 \\
158&$^\mathrm{c}$ & J131659.37+035319.8 & 0.045 & 199 & 5.84 & 226 \\
163&$^\mathrm{c}$ & J132428.24+044629.6 & 0.021 & 91 & 5.81 & 226 \\
174& & J140829.26+562823.5 & 0.133 & 626 & 6.24 & 68 \\
203& & J155909.62+350147.4 & 0.031 & 136 & 6.31 & 991
\enddata
\tablecomments{Basic sample properties, including identification from
\citet{gh07}, SDSS name, redshift, luminosity distance in Mpc,
logarithmic black hole mass in solar units, and time in units of
days between radio and X-ray observations if both exist. Only
sources with an entry for $\Delta T_\mathrm{obs}$ have both X-ray
(new or archival) and radio data for analysis with the fundamental
plane, including GH 140 and GH 158, which have only upper limits in
X-ray, and GH 174, which has only an upper limit in radio. C-sample
sources (Section \ref{sample}) are identified with a superscript c.}
\kgtabend{sample}{Summary of Small AGN Sample}
\subsection{X-Ray Observations}
\label{xrayobs}
We used five archival and six new \emph{Chandra} observations to measure
X-ray luminosities. For new observations, exposure times were 15 ksec
except for the dimmest, GH 140, which had an exposure time of 24 ksec.
New \emph{Chandra} observations were observed on the S3 chip of the
ACIS-S detector. We re-reduced and re-analyzed the archival data.
Data reduction followed the normal pipeline with the most recent
\emph{Chandra} data reduction software package (CIAO version 4.5) and
calibration databases (CALDB version 4.5.5). Source regions were
circles with radii of 4 or 5 pixels centered on the brightest putative
point source consistent with the center of the host galaxy. Given the
distances to the host galaxies (Table \ref{t:sample}), X-ray binaries are not
luminous enough to be a significant source of contamination.
Background regions were annuli with inner radii equal to the
source region radius and outer radii equal to $\sim30$ pixels. We used the
specextract tool to create response matrix and ancillary
response files and extract source and
background spectra.
Spectral fitting was done with the most recent version of Xspec
\citep[v12.6.0q;][]{1996ASPC..101...17A}. Since we were most
interested in the unscattered, intrinsic 2--10 keV flux, we always
included a power-law component in our spectral model with Galactic
absorption set to the value toward each source
\citep{2005A&A...440..775K}. For sources with enough counts, we also
included redshifted intrinsic absorption. In all cases,
as long as the source was detected and total absorption was below $N_\mathrm{H}=10^{20}\ \units{cm^{-2}}$, the inferred intrinsic 2--10 keV flux
from the power-law component was robust. For sources with low
count rates or for which only an upper limit could be determined for
the X-ray flux, we held the photon index constant at $\Gamma=1.7$.
Spectral fitting was done with $C$-stat statistics
\citep{1979ApJ...228..939C}. We report 2--10 keV flux and
$\Gamma$ for each source in Table \ref{t:xrayobs}. Sources GH 119,
and 140 were nondetections; a point source at the location of GH 158
was detected with more than 12 net counts in the 0.5--10 keV band, but
it was not enough to constrain the hard flux at the $3\sigma$ level.
Source 47 had a 2 ksec archival \emph{Chandra} exposure that showed it
likely a Seyfert 2. A 23-ksec \emph{XMM-Newton} archival
observation with a high count rate has been looked at in detail by
\citet{reneeinprep}. They find that the spectrum of GH 47 is well
fitted by a typical type II AGN spectrum consisting of a highly
absorbed power-law component, soft, scattered power-law component, and
a distant reflection component with prominent narrow Fe~K$\alpha$
line. The unabsorbed 2--10 keV flux of the absorbed power-law
component is $2.05\times 10^{-12}\ \units{erg\ cm^{-2}\ s^{-1}}$. We
use the \emph{XMM-Newton} measurement for all subsequent analysis.
As a self-consistency check, we calculated Eddington fractions
($f_\mathrm{Edd}$) of each source assuming the bolometric correction
due to \cite{2004MNRAS.351..169M}. Because the X-ray bolometric
correction depends on the bolometric luminosity, we use the bolometric
luminosity reported in \citet{gh07}. Although using $f_\mathrm{Edd}$
based on optical emission to estimate $f_\mathrm{Edd}$ based on X-ray
emission is circular, it serves as a check for self-consistency. Using
a constant bolometric correction of 20 instead of our adopted circular
method does not make a large difference. We plot the X-ray
$f_\mathrm{Edd}$ as a function of the optically determined
$f_\mathrm{Edd}$ in Figure\ \ref{f:fedd}. The agreement between the two
estimates shows self-consistency.
\kgtabbeg{rrrrr@{\extracolsep{0pt}}ll}
\footnotesize
\tablecaption{X-ray Observations}
\tablewidth{\columnwidth}
\tablehead{
\colhead{GH ID} &
\colhead{Obsid} &
\colhead{MJD} &
\colhead{$t_\mathrm{exp}$} &
\multicolumn{2}{c}{$\log(F_{X}/\units{erg\,s^{-1}\,cm^{-2}})$} &
\colhead{$\Gamma$}
}
\startdata
47 & 504102001 & 54408 & 23 & &$-11.69^{+0.08}_{-0.04}$ & $2.0\pm{0.2}$\\
69 & \dataset[ADS/Sa.CXO#Obs/13858]{13858} & 56097 & 15 & &$-12.43\pm{0.03}$ & $1.9\pm{0.1}$\\
87 & \dataset[ADS/Sa.CXO#Obs/13859]{13859} & 56214 & 15 & &$-12.63\pm{0.04}$ & $2.0\pm{0.2}$\\
106 & \dataset[ADS/Sa.CXO#Obs/11456]{11456} & 55424 & 2 & &$-12.17^{+0.05}_{-0.08}$ & $1.6^{+0.2}_{-0.1}$\\
119 & \dataset[ADS/Sa.CXO#Obs/09234]{9234} & 54551 & 5 & $<$&$-14.39$ & 1.7\\
140 & \dataset[ADS/Sa.CXO#Obs/13860]{13860} & 56200 & 24 & $<$&$-14.05$ & 1.7\\
146 & \dataset[ADS/Sa.CXO#Obs/5664]{5664} & 53428 & 5 & &$-12.94^{+0.14}_{-0.09}$ & 1.7\\
158 & \dataset[ADS/Sa.CXO#Obs/13861]{13861} & 56050 & 15 & $<$&$-12.88$ & 1.7\\
163 & \dataset[ADS/Sa.CXO#Obs/13862]{13862} & 56050 & 15 & &$-12.80\pm{0.07}$ & $1.4\pm{0.2}$\\
174 & \dataset[ADS/Sa.CXO#Obs/13863]{13863} & 56208 & 15 & &$-12.58\pm{0.03}$ & $1.8\pm{0.1}$\\
203 & \dataset[ADS/Sa.CXO#Obs/11479]{11479} & 55234 & 2 & &$-11.62\pm{0.03}$ & $2.3\pm{0.1}$
\enddata
\tablecomments{We provide Obsid, MJD, and exposure time of each
\emph{Chandra} observation in ksec along with 2--10\,keV flux, and
photon index ($\Gamma$) with their 1$\sigma$ uncertainties. When
only an upper limit could be determined, we list the $3\sigma$ upper
limit. Values of $\Gamma$ without uncertainties were fixed at 1.7.
Sources GH 106, GH 119, GH 146, and GH 203 are archival data sets
originally analyzed by \citet{2012ApJ...761...73D},
\citet{2014ApJ...782...55Y}, \citet{2007ApJ...656...84G}, and
\citet{2012ApJ...761...73D}, respectively. Values for source 47 are
from \emph{XMM-Newton} observations with results due to
\citet{reneeinprep}.}
\kgtabend{xrayobs}{X-ray observations}
\kgfigbeg{fedd}
\includegraphics[width=0.95\columnwidth]{./fedd}
\caption{Comparison of $f_\mathrm{Edd}$ estimates. The abscissa shows
$f_\mathrm{Edd}$ determined by \citet{gh07} from optical spectra.
The ordinate shows $f_\mathrm{Edd}$ determined from 2--10 keV
luminosities and bolometric corrections due to
\citet{2004MNRAS.351..169M}. Since the X-ray bolometric corrections
depend on the bolometric luminosity, we use $L_\mathrm{bol}$ as
determined by \citet{gh07}, which serves as a check for
self-consistency. Open circles indicate c-sample sources. The
error bars come from the uncertainty on the X-ray flux only, and the
line shows equality. There is generally good agreement, and the
points above $f_\mathrm{Edd}=1$ are consistent with sub-Eddington
accretion when uncertainties in black hole mass and bolometric
correction are taken into account.}
\kgfigend{fedd}{Eddington fraction}
\subsection{Radio Observations}
\label{radioobs}
Radio data presented in this Letter come from new VLA observations,
taken at 8.4 GHz with 2 GHz bandwidth in the A configuration. The
data were from two programs targeting low-mass AGNs, one selecting
from those with VLA FIRST detections, one from those without.
All observations began with a scan on a flux calibrator. 3C 286 was
used for all observations, except GH 57 and GH 69, for which 3C 147
was used. The flux
calibrator was followed by a phase calibrator (Table
\ref{t:radioobs}). Then we cycled between the source and the phase calibrator
for the remaining duration of the scan.
Flagging and reduction of VLA data followed the standard pipeline
using CASA version 4.0.1. We calibrated fluxes based on the most
recent calibration models, implemented phase corrections, and then
calibrated the bandpass. We averaged the data in bins in time by 30 s
and in frequency by eight channels. Each frequency was converted into an
image and processed with the CLEAN algorithm separately with a maximum
of 5000 iterations, a gain of 0.1, and natural weighting. The
processing used the full width of the $512\times512$ image at a
resolution of 0\farcs05 and typically stopped at a threshold of 0.05
mJy.
The resulting processed images were then inspected for emission at the
location of the AGN. In 10 of the 12 sources, there was an
unambiguous, unresolved point source at the coordinates of the galaxy
center. We attribute this flux to the AGN. For these detected
sources, we calculated the flux density by fitting a two-dimensional
Gaussian to the point source in a $20\times20$ pixel box and using the
total flux reported by the imfit tool. Uncertainties in flux were
calculated as the quadrature sum of uncertainty in the fit and rms
noise of the image. All detections were highly significant
($>10\sigma$). The final flux densities we report (Table
\ref{t:radioobs}) are from the channel centered at 8.5 GHz with an
additional 5\% uncertainty to account for absolute flux calibration.
Non-detections are reported as $3\sigma$ upper limits of the rms
noise of the image.
Detection of the sources in the VLA FIRST survey was not a good
predictor of AGN radio emission. The two non-detections are detected
in VLA FIRST, indicating that the 1.4 GHz VLA FIRST detection is not
associated with AGN activity. All sources without detection at 1.4
GHz were detected at 8.5 GHz. For the detected sources, we attempted
to constrain the spectral index, $\alpha$ ($S_\nu\propto
\nu^{-\alpha}$), by fitting a power law to the fluxes across the
entire bandpass of the radio observations, but the relatively large
uncertainties meant that we only had a weak constraint. All
measurements were consistent with $\alpha=0.7$, which we assume for
subsequent analysis.
\kgtabstarbeg{rlrlllr@{\extracolsep{0pt}}ll}
\footnotesize
\tablecaption{VLA Observations}
\tablewidth{\textwidth}
\tablehead{
\colhead{GH ID} &
\colhead{SB ID} &
\colhead{MJD} &
\colhead{Cal.} &
\colhead{$t_\mathrm{exp}$} &
\colhead{Beam Size} &
\multicolumn{2}{c}{$S_\nu/\mathrm{mJy}$} &
\colhead{$\mathrm{rms}/\mathrm{mJy}$}
}
\startdata
47 & 12097202 & 56298 & J0830+2410 & 59.5 & $0\farcs37\times0\farcs28$ & &$0.93 \pm{0.05}$ & 0.014 \\
69 & 12452868 & 56274 & J0914+0245 & 34.7 & $0\farcs36\times0\farcs26$ & &$0.60 \pm{0.03}$ & 0.015\\
87 & 12465022 & 56274 & J1008+0730 & 33.2 & $0\farcs42\times0\farcs26$ & &$0.42 \pm{0.03}$ & 0.022\\
94 & 12469924 & 56274 & J0958+6533 & 33.9 & $0\farcs32\times0\farcs25$ & $<$&$0.29$ & 0.087\\
101 & 12118330 & 56243 & J1033+6051 & 57.6 & $0\farcs33\times0\farcs25$ & &$0.54 \pm{0.03}$ & 0.012\\
106 & 12117646 & 56246 & J1110+6028 & 57.6 & $0\farcs31\times0\farcs25$ & &$1.10 \pm{0.06}$ & 0.014\\
140 & 12465449 & 56276 & J1217+5835 & 33.8 & $0\farcs28\times0\farcs24$ & &$0.37 \pm{0.03}$ & 0.018\\
146 & 11470757 & 56252 & J1229+0203 & 56.3 & $0\farcs42\times0\farcs26$ & &$0.45 \pm{0.03}$ & 0.017\\
158 & 12465639 & 56276 & J1256$-$0547 & 30.9 & $0\farcs30\times0\farcs24$ & &$0.57 \pm{0.03}$ & 0.023\\
163 & 12465813 & 56277 & J1347+1217 & 32.1 & $0\farcs49\times0\farcs26$ & &$0.50 \pm{0.03}$ & 0.022\\
174 & 12469748 & 56276 & J1419+5423 & 34.1 & $0\farcs37\times0\farcs24$ & $<$&$0.10$ & 0.028\\
203 & 12118194 & 56225 & J1602+3326 & 57.4 & $0\farcs27\times0\farcs26$ & &$0.58 \pm{0.04}$ & 0.015
\enddata
\tablecomments{VLA scheduling block identification, MJD
of observation, gain calibrator, time on source in minutes, size of synthesized beam, flux density of source, and rms of map in mJy. All observations used 3C286 for
flux calibration and bandpass calibration except for GH 47 and GH
69, which used 3C147. The flux density uncertainty includes a
5\% uncertainty in absolute flux calibration, which dominates the
total uncertainty. Upper limits are $3\sigma$
values.}
\kgtabstarend{radioobs}{VLA observations}
\section{Analysis and Discussion}
\label{analysis}
\label{discussion}
Our primary analysis is to compare these low-mass AGNs with the earlier
sample of SMBHs with dynamical masses on the fundamental plane. As
mentioned in Section \ref{sample}, we update the \citet{gh07} masses
using the most recent H$\alpha$ AGN-mass scaling relations and adopt
conservative uncertainties of 0.5 dex. The uncertainty in mass is
much larger than the statistical uncertainties in the H$\alpha$
luminosity and line-width measurements, the uncertainties in the
best-fit AGN-mass scaling relations, and the intrinsic scatter in the
scaling relations. The larger uncertainty, however, should encompass
systematic uncertainties in mass estimation such as extrapolation of
scaling relations to low masses \citep[e.g.,][]{2010ApJ...721...26G},
linking H$\beta$ scaling relations to H$\alpha$ scaling relations, and
imperfect decomposition of narrow lines and broad lines, which may be
especially difficult for narrow-line Seyfert 1 AGNs
\citep{2009ApJ...692..246D}.
We calculate 5 GHz radio luminosities from our measured 8.5 GHz flux
densities assuming a spectral index of $\alpha=0.7$. Note that $L_R\equiv\nu L_\nu$, whereas the 2--10 keV X-ray luminosity, $L_X$, is
a bandpass luminosity.
From the 13 sources in Table \ref{t:sample}, we have usable data in
both X-ray and radio for 10, including 3 sources with upper limits on
one of the two quantities. We compare these data to the two
fundamental-plane fits in Figure \ref{f:fpview} and find that they are
better predicted by the universal (all black holes) fit. The low-mass
AGNs are inconsistent with the SMBH-only relation, having higher $L_R$,
lower $L_X$, and/or lower $M$ than predicted. Compared to the
universal relation, the low-mass AGNs are within the scatter.
\kgfigstarbeg{fpview}
\includegraphics[width=0.45\textwidth]{./fpview_r_mx_smbh}
\includegraphics[width=0.45\textwidth]{./fpview_r_mx_full}\\
\includegraphics[width=0.45\textwidth]{./fpview_m_rx_smbh}
\includegraphics[width=0.45\textwidth]{./fpview_m_rx_full}
\caption{New data for the current sample of small AGNs (red circles) and
for SMBHs with dynamically determined black hole masses (black
crosses) due to \citet{2009ApJ...706..404G}. Open circles indicate
sources from the \citet{gh07} c-sample. Top panels have $L_R$ as
the dependent variable, and bottom panels have $M$ as the dependent
variable. The left panels are projected to display the edge-on view
of the best fit to SMBHs with dynamical masses only, and the right
panels for both stellar-mass and SMBHs with
dynamical masses. Fits do not include low-mass AGNs, and thus we may compare the new, low-mass data to the values predicted by the fitted relations. Low-mass AGNs
are inconsistent with the prediction of the left panels and
consistent with the prediction of the right panels.}
\kgfigstarend{fpview}{Fundamental plane projections}
The low-mass AGNs are slightly but systematically offset from the
better-fitting SMBH-only plane in one direction (Figure \ref{f:fpview},
right panel). The median deviations in the $L_R$, $M$, and $L_X$
directions are 0.63, $-0.81$, and $-0.95$ dex, respectively. The
magnitude of these deviations is within the scatter, but of seven sources
with both X-ray and radio detections, six lie to one side of the plane.
If an individual source is equally likely to lie on either side of the
plane, the probability of six or more out of seven lying on the same side
(high or low, i.e., two-sided) by chance is 0.125. If we
exclude the c-sample source GH 163 from this calculation, then chance
alignment for five of six sources would happen 0.219 of the time. This is
only weak evidence of a systematic deviation, but we briefly
speculate on potential reasons if this trend were to continue with
more data. Despite this speculation, below we conclude that the
low-mass AGNs are consistent with the full fundamental plane, which we
argue should be considered the better model.
Because black hole mass is a notoriously difficult quantity to
estimate and is the only indirect quantity considered, we give it
special attention. The \citet{gh07} sample is selected for broad
H$\alpha$ lines with small FWHM. It is
possible, but not certain, that broad emission lines come from the
base of a disk wind \citep[e.g.,][]{2013A&A...551L...6K} with
predominantly rotational motions. The resulting FWHM measurement
would be subject to orientation effects that could lead to a
preferential selection of face-on inclinations and underestimation of
the true mass. This possibility is supported by the LAMP 2008 campaign's
finding that lower inclination objects return larger virial
coefficients \citep{2013arXiv1311.6475P}.
Several other lines of evidence, however, point to, on average,
accurate mass estimation. First, reverberation mapping of GH 126
indicates a small mass in agreement with the H$\alpha$ single-epoch
estimate \citep{2011ApJ...741...66R}. Second, the expected black hole
mass based on host galaxy stellar velocity dispersion generally agrees
with the single-epoch estimates \citep{2011ApJ...739...28X}, though
megamaser measurements of black hole mass in small galaxies show that
the masses can be much smaller than predicted by velocity dispersion
\citep{2010ApJ...721...26G} and low-mass AGNs appear undermassive
relative to their bulge luminosity \citep{2011ApJ...737L..45J}. Given
that many of the host galaxies are likely to be pseudobulges
\citep{2011ApJ...742...68J}, black hole mass--host galaxy property
scaling relations may not apply to the majority of these galaxies
\citep{2011Natur.469..374K, 2010ApJ...721...26G}. Finally,
\citet{reneeinprep} recently found that the X-ray variability of a
different sample of \citet{gh07} AGNs is as expected for low-mass
sources. Given the above evidence, we find it unlikely that
underestimated masses are the cause of the offset.
Another possibility is that the offset results from different
accretion--jet interactions. The data are not conclusive, but X-ray
properties of AGNs may depend on black hole mass
\citep{2009ApJ...698.1515D, 2012ApJ...761...73D}. Given the inferred
near-ultraviolet luminosity of these low-mass AGNs, their spectral
index between 2500 \AA\ and 2 keV ($\alpha_{OX}$) may be flatter than
expected, and their 2 keV X-ray emission sometimes weaker than
expected, though selection effects may complicate the issue. This may
be evidence of slim-disk accretion, lack of a Comptonizing corona, or
high intrinsic absorption. Thus the 2 keV X-ray emission of low-mass
AGNs may not be as reliable a probe of accretion power as it is in
higher mass AGNs. If the difference were a result of a lack of a
Comptonizing corona, then this would lead to an absence of 2--10 keV
flux, which we assume is from inverse Compton scattering. Rather, we
observe X-ray flux levels that are consistent with expectations based
on the accretion rate and bolometric corrections (Figure \ref{f:fedd}).
If the difference in 2 keV luminosity is, instead, a result of high
intrinsic absorption of any form, our conclusions should be unaffected
owing to our use of hard flux and deeper observations.
Based on the relatively small offset, the reliability of 2--10 keV
flux measurements in the face of absorption, and the low significance
of a possible offset of the low-mass points from the fundamental
plane, we conclude that low-mass AGNs do in fact follow the full
$M$--$L_R$--$L_X$ plane. Our results suggest that stellar-mass black
holes and supermassive black holes follow the same relation. A
similar test using the most massive black holes would provide
additional support.
Given the low mass and relatively high accretion rates here, we can
now state with confidence that (1) it is appropriate to use the
fundamental plane to estimate SMBH masses smaller
than $\sim10^7\ \msun$ \citep[e.g.,][]{2011Natur.470...66R}, (2) it
is possible to use the fundamental plane to test for intermediate mass
black holes \citep[e.g.,][]{2013MNRAS.436.1546M}, and (3) it is
possible to use the fundamental plane to estimate black hole masses at
high accretion rates \citep[e.g.,][]{2011ApJ...738L..13M}.
\hypertarget{ackbkmk}{}%
\acknowledgements
\bookmark[level=0,dest=ackbkmk]{Acknowledgments}
We thank Jenny Greene for a valuable referee report and Elena Gallo
for helpful comments. K.G.\ acknowledges support provided by NASA
through Chandra Award G02-13111X (\emph{Chandra X-Ray Observatory} operated by NASA under contract
NAS8-03060). Results reported here are based on new and archival
\emph{Chandra} data. We used NED, ADS, and CIAO and are grateful.
\bibliographystyle{apjads}
\hypertarget{refbkmk}{}%
\bookmark[level=0,dest=refbkmk]{References}
|
\section{Introduction}
\label{sec:intro}
The basis of many important problems in genetics is to find an expression for a sampling distribution or likelihood. Valuable tools in this endeavour are stochastic models of allele frequency evolution forwards in time, and their dual genealogical processes backwards in time. In particular, the numerous variants of the Wright-Fisher diffusion and Kingman's coalescent, respectively, have focused attention on the scaling limit as the population size goes to infinity, leading from a (complicated) finite-population model of reproduction to a (simpler) infinite-population limit. At a single genetic locus, the problem of computing sampling distributions in these models is well studied, with even some closed-form formulas available \citep{wri:1949, ewe:1972,jen:son:2011,bha:etal:2012}. However, with ongoing technological developments in high-throughput DNA sequencing, large genomic datasets are becoming available and it is necessary to consider multilocus models. Inter-locus recombination quickly makes such models intractable; for neither the Wright-Fisher diffusion with recombination nor the coalescent with recombination---or \emph{ancestral recombination graph} (ARG)--
is it possible to obtain a closed-form expression for the sampling distribution. This has remained a notoriously difficult problem, and to make progress using these models it has usually been necessary to resort to computationally-intensive techniques such as importance sampling \citep{gri:mar:1996, fea:don:2001, gri:etal:2008, jen:gri:2011}, Markov chain Monte Carlo \citep{kuh:etal:2000, nie:2000, wan:ran:2008, ras:etal:2014}, or other numerical approximations
\citep{boi:loi:2007, miu:2011}. Denoting the population-scaled recombination parameter by $\rho$, only in the special cases of $\rho = 0$ or $\rho = \infty$ is it possible to make progress analytically, since then we are back to a single locus, or to many independent single loci, respectively.
In another direction, we
have considered an analytic approach to the problem, as follows. Denote the observed sample configuration at two loci by ${\bfmath{n}}$ and its sampling probability by $q({\bfmath{n}}; \rho)$ (to be defined precisely below). Consider the asymptotic expansion in inverse powers of $\rho$:
\begin{equation}
\label{eq:main}
q({\bfmath{n}}; \rho) = q_0({\bfmath{n}}) + \frac{q_1({\bfmath{n}})}{\rho} + \frac{q_2({\bfmath{n}})}{\rho^2} + \cdots,
\end{equation}
where for convenience we suppress the dependence of these terms on other parameters of the model. Under an infinite-alleles type of mutation, we
obtained closed-form formulas for $q_0({\bfmath{n}})$ and $q_1({\bfmath{n}})$ in terms of the marginal \emph{one}-locus sampling probabilities, and a decomposition of $q_2({\bfmath{n}})$ into a closed-form term plus a second part which is evaluated easily by dynamic programming \citep{jen:son:2010:AAP}. (The result is stated more precisely in \thmref{thm:jen:son:2009} below.) This provides the first closed-form extension of Ewens' Sampling Formula \citep{ewe:1972} to handle finite amounts of recombination. It has been extended subsequently to include more general models of mutation \citep{jen:son:2009:G}, natural selection \citep{jen:son:2012:AAP}, higher-order terms \citep{jen:son:2012:AAP}, and more than two loci \citep{bha:son:2012}, and has had practical implications for genomic inference \citep{cha:etal:2012}. One particularly appealing conclusion of these works is that both $q_0({\bfmath{n}})$ and $q_1({\bfmath{n}})$ are \emph{universal}; that is, their functional form is invariant to our assumptions about mutation and selection acting marginally at each locus. The effects of these marginal processes are entirely subsumed into the relevant one-locus sampling distributions.
The simple and universal forms for $q_0({\bfmath{n}})$ and $q_1({\bfmath{n}})$ provide strong circumstantial evidence that there exists an underlying stochastic process which is much simpler than the standard models for finite amounts of recombination. In particular, we previously conjectured \citep{jen:son:2010:AAP} the existence of a process which is both much simpler than the standard models based on the Wright-Fisher diffusion or on the ARG, and is in agreement with the sampling distribution \eref{eq:main} up to $O(\rho^{-2})$. The goal of this paper is to describe such a process. In fact, using different arguments we describe two such processes, obtaining both a limiting diffusion and a coalescent process with these properties. In the diffusion approximation, the key idea is to suppose that the probability $r$ of a recombination per individual per generation scales as $N^{-\beta}$ as the population size $N\to\infty$, for $0 < \beta < 1$, rather than the usual choice of $\beta = 1$. Interest in asymptotically large recombination rates is reasonable because of extensive recombination rate heterogeneity along chromosomes in e.g.\ humans, strong recombination rates in some species such as \emph{Drosophila melanogaster} \citep{cha:etal:2012}, and because of the need to understand the long-range dependencies between well-separated loci. Our diffusion in this scaling is intimately related to the central limit theorem for density dependent population processes \citep[see][Theorem 11.2.3]{eth:kur:1986}, which has been analyzed in genetics---for models of strong mutation rather than strong recombination---by \citet{fel:1951} and \citet{nor:1975:SIAM}.
A closely related scaling in the context of $\Xi$-coalescent processes was also recently explored by \citet{bir:etal:2013} (in that paper $\beta = 1$ but with timescale $N^2$). The coalescent approach, meanwhile, uses a coupling argument. Intuitively, we would like to couple the ARG to the limiting case of two independent coalescent trees ($\rho = \infty$). To account for contributions to the sampling distribution of $O(\rho^{-1})$, we must quantify the ``leading order reasons'' for such a coupling to fail. When $\rho$ is large but finite, lineages in the ARG ancestral to both loci undergo recombination backwards in time very rapidly, until the first time $U$ that no such lineage survives. In this paper we show that, roughly speaking, in order to recover the sampling distribution up to $O(\rho^{-1})$ we need consider only the following type of exceptional event: \emph{a coalescence occurs more recently than time $U$ in the ARG, and the coalescence is between two lineages each of which is ancestral to both of the two loci}. This observation enables us to \emph{define} a simple coalescent process which allows for at most one of these events but is otherwise very similar to the easy limiting process corresponding to $\rho = \infty$.
The paper is organized as follows. In \sref{sec:notation} we specify our notation and summarize previous research. Novel diffusion and coalescent processes are introduced in Sections \ref{sec:diffusion} and \ref{sec:coalescent}, respectively, and we conclude in \sref{sec:discussion} with a brief discussion.
\section{Notation and previous results}
\label{sec:notation}
For $M \in \mathbb{N} = \{0,1,2,\ldots\}$, let $[M] := \{1,2,\ldots, M\}$. The complement of a set $J$ is written $J^\complement$. Denote the Kronecker delta by $\delta_{ij}$ which takes the value 1 if $i = j$ and 0 otherwise. Let ${\bfmath{e}}_i$ denote a unit vector whose $j$th entry is $\delta_{ij}$, and let ${\bfmath{e}}_{ij}$ denote a matrix with $(k,l)$th entry equal to $\delta_{ik}\delta_{jl}$.
For a vector ${\bfmath{v}} \in \mathbb{R}^d$ we denote by $|{\bfmath{v}}|$ the usual Euclidean norm.
Denote the $k \times l$ zero matrix by $\bfmath{0}_{k\times l}$ and the $k \times k$ identity matrix by $\bfmath{I}_k$. We will replace a subscript with a ``$\cdot$'' to denote summation over that index. A prime symbol $'$ will denote vector or matrix transpose. For $z\in \mathbb{R}_{\geq 0}$ and $n \in \mathbb{N}$, $\phammer{z}{n} := z(z+1)\cdots(z+n-1)$ denotes the $n$th ascending factorial of $z$. Finally, for a matrix $\bfmath{R}$ of processes we let $[\bfmath{R}]_t = ([R_i,R_j]_t)_{i,j}$ denote the matrix of corresponding covariation processes.
Consider the usual diffusion limit of an exchangeable model of random mating with constant population size of $N$ haplotypes.
Our interest will be in a sample from this population at two loci, which we call A and B, with the probability of mutation per haplotype per generation denoted by $u_A$ and $u_B$ respectively. In the diffusion limit we let $N\to\infty$ and $u_A$, $u_B \to 0$ while the population-scaled parameters $\theta_A = 2Nu_A$ and $\theta_B = 2Nu_B$ remain fixed. In this paper we will suppose a \emph{finite-alleles} model of mutation such that a mutation to an allele $i$ in type space $E_A = [K]$, $K \in \mathbb{N}$, takes it to allele $k \in [K]$ with probability $P_{ik}^A$, with $E_B = [L]$ and $P_{jl}^B$, $j,l \in [L]$ defined analogously. (As we discover below, the mutation model is not important and we could pose something more complicated with little extra effort.)
The probability of a recombination between the two loci per haplotype per generation is denoted by $r$, and we assume that $\rho_\beta = 2N^\beta r$ is fixed as $N \to \infty$, for some fixed $\beta \in (0,1]$. Previous work has focused on the case $\beta = 1$ with time measured in units of $N$ generations. For consistency with the usual notation we write $\rho = \rho_1$.
A sample from this model comprises $a$ haplotypes observed only at locus A, $b$ haplotypes observed only at locus B, and $c$ haplotypes observed at both loci. The sample configuration is denoted by ${\bfmath{n}} = ({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$ where ${\bfmath{a}} = (a_i)_{i\in [K]}$ and $a_i$ is the number of haplotypes observed to exhibit allele $i$ at locus A; ${\bfmath{b}} = (b_j)_{j\in [L]}$ where $b_j$ is the number of haplotypes observed to exhibit allele $j$ at locus B; and ${\bfmath{c}} = (c_{ij})_{i\in [K], j\in [L]}$ where $c_{ij}$ is the number of haplotypes with allele $i$ at locus A and allele $j$ at locus B. Thus,
\[
\begin{array}{ccc}
a = \displaystyle\sum_{i=1}^K a_i, & b = \displaystyle\sum_{j=1}^L b_j, & c = \displaystyle\sum_{i=1}^K\sum_{j=1}^L c_{ij},
\end{array}
\]
and we let $n = a + b + c$. We further write ${\bfmath{c}}_A = (c_{i\cdot})_{i\in[K]}$ and ${\bfmath{c}}_B = (c_{\cdot j})_{j\in[L]}$ to denote the marginal sample configurations of ${\bfmath{c}}$ restricted to locus A and locus B respectively. Finally, we use $q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$ to denote the probability that when we sample $n$ haplotypes in some order from the population at stationarity we obtain the unordered configuration $({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$; by sampling exchangeability this is indeed a function only of the unordered configuration $({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$. For convenience we suppress the dependence of this quantity on the model parameters and on $\beta$. The main result motivating this work is an expansion for $q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$ for the case of $\beta=1$, and later we will show that this expansion holds for all $\beta \in (0,1]$.
\begin{theorem}[See \citet{jen:son:2009:G}]
\label{thm:jen:son:2009}
Consider the following asymptotic expansion for $q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$ under the diffusion limit with $\beta = 1$:
\[
q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = q_0({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) + \frac{q_1({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})}{\rho} + O\left(\frac{1}{\rho^2}\right), \quad \text{as }\rho\to\infty,
\]
with $q_0$, $q_1$, $\ldots$ independent of $\rho$. Then the zeroth order term is given by
\begin{equation}
q_0({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = q^A({\bfmath{a}} + {\bfmath{c}}_A)q^B({\bfmath{b}} + {\bfmath{c}}_B),
\label{eq:zerothorder}
\end{equation}
and the first order term is given by
\begin{align}
\label{eq:firstorder}
q_1({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = {}& \binom{c}{2}q^A({\bfmath{a}}+{\bfmath{c}}_A)q^B({\bfmath{b}}+{\bfmath{c}}_B) \nonumber\\
{}& -q^B({\bfmath{b}}+{\bfmath{c}}_B)\sum_{i=1}^K \binom{c_{i\cdot}}{2}q^A({\bfmath{a}}+{\bfmath{c}}_A-{\bfmath{e}}_i) \nonumber\\
{}& -q^A({\bfmath{a}}+{\bfmath{c}}_A)\sum_{j=1}^L \binom{c_{\cdot j}}{2}q^B({\bfmath{b}}+{\bfmath{c}}_B-{\bfmath{e}}_j) \nonumber\\
{}& +\sum_{i=1}^K \sum_{j=1}^L \binom{c_{ij}}{2}q^A({\bfmath{a}}+{\bfmath{c}}_A-{\bfmath{e}}_i)q^B({\bfmath{b}}+{\bfmath{c}}_B-{\bfmath{e}}_j),
\end{align}
where $q^A$, $q^B$ are the marginal sampling distributions at locus A and locus B, respectively.
\end{theorem}
\begin{remark} \label{rem:jen:son:2009}
Under a neutral, finite-alleles model of mutation, if mutation is \emph{parent independent}---that is, $P_{ki}^A = P_{i}^A$, $i,k \in [K]$, and $P_{lj}^B = P_{j}^B$, $j,l \in [L]$, then $q^A({\bfmath{a}})$ and $q^B({\bfmath{b}})$ are known in closed-form:
\[
\begin{array}{lcr}
q^A({\bfmath{a}}) = \displaystyle\frac{1}{\phammer{\theta_A}{a}}\prod_{i=1}^K \phammer{\theta_A P^A_i}{a_i}, & \text{ and } &
q^B({\bfmath{b}}) = \displaystyle\frac{1}{\phammer{\theta_B}{b}}\prod_{j=1}^L \phammer{\theta_B P^B_j}{b_j}.
\end{array}
\]
These expressions follow, for example, from the moments of the Wright-Fisher diffusion with parent-independent mutation, whose stationary distribution at locus A is $\text{Dirichlet}(\theta_\A P_1^A,\ldots, \theta_\A P_{K-1}^A)$ \citep{wri:1949}, and similarly at locus B.
\end{remark}
\begin{remark}
The zeroth-order decomposition is well known \citep[e.g.][]{eth:1979:JAP} and also intuitive, since the two loci become independent as $\rho\to\infty$.
\end{remark}
\thmref{thm:jen:son:2009} can be obtained by diffusion \citep{jen:son:2012:AAP} or by coalescent \citep{jen:son:2009:G, jen:son:2010:AAP} arguments. In this paper we address both approaches in further detail.
\section{Diffusion model}
\label{sec:diffusion}
In this section we extend the above results by obtaining a full description of a simple diffusion process such that its sampling distribution is known \emph{exactly} and has a Taylor expansion about $\rho=\infty$ consistent with \eref{eq:zerothorder} and \eref{eq:firstorder}. For simplicity we will obtain our diffusion as the limit of an appropriately rescaled Moran model, although we expect our results to hold for a more general class of discrete models of reproduction within the domain of convergence of the Wright-Fisher diffusion.
\subsection{Neutral Moran model}
A population of $N$ haploid, monoecious individuals evolves as a multitype birth-and-death process in continuous time. Each individual carries a haplotype comprising a pair of alleles $(i,j) \in [K] \times [L]$, one at locus A and one at locus B. Let $Z_{ij}(\uptau) \in \{0,1,\ldots,N\}$ denote the number of $(i,j)$ haplotypes in the population at time $\uptau\in \mathbb{R}_{\geq 0}$, and $\bfmath{Z}(\uptau) = (Z_{ij}(\uptau))_{i \in [K], j \in [L]}$. The population evolves as follows. At rate $N^2/2$ a reproduction event occurs, in which an individual is chosen uniformly at random from the population to die. It is replaced by a copy of another individual also chosen uniformly at random (the same individual could be chosen; whether sampling is with or without replacement does not affect the diffusion limit). Independently, each locus of each haplotype undergoes mutation: any locus A mutates at rate $\theta_\A/2$ and its allele is updated according to the transition matrix $\bfmath{P}^A = (P_{ik}^A)_{i,k\in[K]}$; similarly any locus B mutates at rate $\theta_\B/2$ and its allele is updated according to $\bfmath{P}^B = (P_{jl}^B)_{j,l\in[L]}$. Finally, each haplotype independently undergoes recombination at rate $\rho/2$: at such an event, it is replaced by a haplotype formed by sampling two alleles (one for each locus) independently from the population. Putting all this together, the rate at which a haplotype $(i,j)$ dies and is replaced by a haplotype $(k,l)$ when $\bfmath{Z}(\uptau) = {\bfmath{z}}$ is given by
\begin{multline*}
\lambda_{ij,kl}^{(N)}({\bfmath{z}}) = \frac{z_{ij}}{N}\left[\frac{N^2}{2}\frac{z_{kl}}{N} + N\left(\frac{\theta_\A}{2}P_{ik}^A\delta_{jl} + \frac{\theta_\B}{2}P_{jl}^B\delta_{ik} + \frac{\rho}{2}\frac{z_{k\cdot}}{N}\frac{z_{\cdot l}}{N}\right)\right],\\ (i,j),(k,l) \in [K]\times [L].
\end{multline*}
Notice that, as is standard \citep[e.g.][]{baa:her:2008}, we decouple the mutation and recombination mechanisms from reproduction (and from each other). This simplifies the analysis without unduly affecting the diffusion limit.
We will change variables by introducing the collection
\[
\bfmath{M}^{(N)}(\uptau) := \{\bfX^{(N)}(\uptau), \bfY^{(N)}(\uptau), \bfD^{(N)}(\uptau)\},
\]
where
\begin{align*}
\bfX^{(N)}(\uptau) = (X^{(N)}_i(\uptau))_{i\in[K]} &= \left(\frac{Z_{i\cdot}(\uptau)}{N}: i \in [K]\right),\\
\bfY^{(N)}(\uptau) = (Y^{(N)}_j(\uptau))_{j\in[L]} &= \left(\frac{Z_{\cdot j}(\uptau)}{N}: j \in [L]\right),\\
\bfD^{(N)}(\uptau) = (D^{(N)}_{ij}(\uptau))_{i\in[K],j\in[L]} &= \left(\!\frac{Z_{ij}(\uptau)}{N} - \frac{Z_{i\cdot}(\uptau)}{N}\frac{Z_{\cdot j}(\uptau)}{N}: i \in [K], j\in[L] \!\right).
\end{align*}
That is, we describe the state of the Moran model at time $\uptau$ by the marginal allele frequencies
and the coefficients of linkage disequilibrium \citep[see, e.g.][p69, p227]{ewe:2004:I}.
We will write this succinctly by arranging the variables in a linear order:
\[
(X_1^{(N)},\ldots, X_K^{(N)}, Y_1^{(N)}, \ldots, Y_L^{(N)}, D^{(N)}_{11}, \ldots, D^{(N)}_{KL})',
\]
and thinking of $\bfmath{M}^{(N)}(\uptau)$ as a vector of length $\Lambda := K + L + KL$. The process $(\bfmath{M}^{(N)}(\uptau) : \uptau = 0,1,\ldots)$ is then Markov on a state space we denote by $\Delta_{KL-1}^{(N)}$, which is a rational subset (those points consistent with $\sum_{i=1}^K\sum_{j=1}^L Z_{ij} = N$) of the $(KL - 1)$-dimensional shifted simplex
\begin{multline*}
\Delta_{KL-1} = \Bigg\{ ({\bfmath{x}},{\bfmath{y}},{\bfmath{d}}) \in [0,1]^{K} \times [0,1]^L \times [-1,1]^{KL} : \\\sum_{i=1}^K x_i = 1 = \sum_{j=1}^L y_j, \sum_{i=1}^K d_{ij} = 0 = \sum_{j=1}^L d_{ij}\Bigg\}.
\end{multline*}
To find the diffusion limit we first need the conditional means and covariances of the increments
\[
\Delta\bfmath{M}^{(N)} (\uptau) := \bfmath{M}^{(N)}(\uptau + \dd\uptau) - \bfmath{M}^{(N)}(\uptau).
\]
From these, and under the assumption that $\theta_\A$, $\theta_\B$, and $\rho$ are fixed as $N\to\infty$, it is possible to show that the model converges to a (Wright-Fisher) diffusion limit \citep[Example 10.3.9, p433]{eth:kur:1986}. Recall however that our interest is when $\rho_\beta$, rather than $\rho$, is fixed, so below we write these increments in terms of $\rho_\beta$ using $\rho = \rho_\beta N^{1-\beta}$.
In the following, for convenience we drop the dependence on $\uptau$.
\begin{proposition}
\label{prop:WF}
In the neutral two-locus Moran model with mutation and recombination, the conditional means and covariances of increments of $\bfmath{M}^{(N)}$
are given by
\allowdisplaybreaks
\begin{align}
\label{eq:Xi}\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\mathbb{E}[\Delta X_i^{(N)}\mid \bfmath{M}^{(N)}] = {}& \frac{\theta_\A}{2}\sum_{k=1}^K (P_{ki}^A - \delta_{ik}
X_k^{(N)}, \\
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\mathbb{E}[\Delta Y_j^{(N)}\mid \bfmath{M}^{(N)}] = {}& \frac{\theta_\B}{2}\sum_{l=1}^L (P_{lj}^B - \delta_{jl}
Y_l^{(N)},\\
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\mathbb{E}[\Delta D_{ij}^{(N)}\mid \bfmath{M}^{(N)}] = {} & -\frac{\rho_\beta}{2N^{\beta-1}}D^{(N)}_{ij} - D^{(N)}_{ij} \notag\\ {}&+ \frac{\theta_\A}{2}\sum_{k=1}^K (P_{ki}^A - \delta_{ik}
D^{(N)}_{kj} \notag\\
{}& {}+ \frac{\theta_\B}{2}\sum_{l=1}^L (P_{lj}^B - \delta_{jl}
D^{(N)}_{il} + O\left(\frac{1}{N^{\beta}}\right),\label{eq:Dij
\end{align}
\begin{align}
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\bbC\textrm{ov}[\Delta X_i^{(N)},\Delta X_k^{(N)}\mid \bfmath{M}^{(N)}] = {}& X^{(N)}_i(\delta_{ik} - X^{(N)}_k) + O\left(\frac{1}{N^\beta}\right), \notag\\%\label{eq:XiXk}\\
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\bbC\textrm{ov}[\Delta Y_j^{(N)},\Delta Y_l^{(N)}\mid \bfmath{M}^{(N)}] = {}& Y^{(N)}_j(\delta_{jl} - Y^{(N)}_l) + O\left(\frac{1}{N^\beta}\right),\notag\\
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\bbC\textrm{ov}[\Delta X_i^{(N)}, \Delta Y_j^{(N)}\mid \bfmath{M}^{(N)}] = {}& D^{(N)}_{ij} + O\left(\frac{1}{N^\beta}\right),\notag\\%\label{eq:XiYj}\\
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\bbC\textrm{ov}[\Delta X_i^{(N)}, \Delta D^{(N)}_{kl} \mid \bfmath{M}^{(N)}] = {}& D^{(N)}_{kl}(\delta_{ik} - X^{(N)}_i) - X^{(N)}_kD^{(N)}_{il} \notag\\ &{}+ O\left(\frac{1}{N^{\beta}}\right), \notag\\%\label{eq:XiDkl}\\
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\bbC\textrm{ov}[\Delta Y_j^{(N)}, \Delta D^{(N)}_{kl} \mid \bfmath{M}^{(N)}] = {}& D^{(N)}_{kl}(\delta_{jl} - Y^{(N)}_j) - Y^{(N)}_lD^{(N)}_{kj} \notag\\ &{}+ O\left(\frac{1}{N^{\beta}}\right),\notag\\
\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\bbC\textrm{ov}[\Delta D_{ij}^{(N)}, \Delta D_{kl}^{(N)} \mid \bfmath{M}^{(N)}] = {}& X_i^{(N)}Y_j^{(N)}(\delta_{ik} - X_k^{(N)})(\delta_{jl} - Y_l^{(N)}) \notag\\
& \hspace{-40pt} {}+ D^{(N)}_{kj}X_i^{(N)}Y_l^{(N)} + D_{il}^{(N)}X_k^{(N)}Y_j^{(N)} \notag\\
& \hspace{-40pt}{}+ D^{(N)}_{ij}(X_k^{(N)}Y_l^{(N)} - \delta_{ik}Y_l^{(N)} - \delta_{jl}X_k^{(N)})\notag\\
& \hspace{-40pt}{}+ D^{(N)}_{kl}(X_i^{(N)}Y^{(N)}_j - \delta_{ik}Y_j^{(N)} - \delta_{jl}X_i^{(N)}) \notag\\
& \hspace{-40pt}{}+ D_{ij}^{(N)}(\delta_{ik}\delta_{jl} - D^{(N)}_{kl})
+ O\left(\frac{1}{N^{\beta}}\right).\nota
\end{align}
Higher order moments of order $m\geq 2$ are $O(N^{-(m-2)})$.
\end{proposition}
\begin{proof}
These expressions follow directly from the first four moments of $\bfmath{Z}(\uptau+\dd\uptau)\mid \bfmath{Z}(\uptau)$, which are easily computed by noting that
\begin{multline*}
\mathbb{E}[f(\bfmath{Z}(\uptau + \dd\uptau))\mid \bfmath{Z}(\uptau) = {\bfmath{z}}] = \sum_{(i,j)}\sum_{(k,l)} f({\bfmath{z}} - {\bfmath{e}}_{ij} + {\bfmath{e}}_{kl})\lambda_{ij,kl}^{(N)}({\bfmath{z}}) \dd\uptau \\
{}+ f({\bfmath{z}})\left[1 - \frac{N}{2}(N + \theta_\A + \theta_\B + \rho)\dd\uptau\right] + o(\dd\uptau).
\end{multline*}
For example, choosing $f({\bfmath{z}}) = z_{uv}$ we find
\begin{multline*}
\mathbb{E}[Z_{uv}(\uptau + \dd\uptau)\mid \bfmath{Z}(\uptau) = {\bfmath{z}}] = z_{uv} \\ {}+ N\left[\frac{\theta_\A}{2}\sum_{k=1}^K(P_{ku}^A - \delta_{ku}) \frac{z_{kv}}{N} + \frac{\theta_\B}{2}\sum_{l=1}^L(P_{lv}^A - \delta_{lv}) \frac{z_{ul}}{N} + \frac{\rho}{2}\left(\frac{z_{u\cdot}}{N}\frac{z_{\cdot v}}{N} - \frac{z_{uv}}{N}\right)\right]\dd\uptau \\{}+ o(\dd\uptau),
\end{multline*}
and hence we recover \eqref{eq:Xi} via
\begin{align*}
\mathbb{E}[\Delta X_u\mid \bfmath{M}^{(N)}] &= \frac{1}{N}\sum_{v=1}^L \left(\mathbb{E}[Z_{uv}(\uptau + \dd\uptau)\mid \bfmath{Z}(\uptau)] - Z_{uv}\right) \\ &= \frac{\theta_\A}{2}\sum_{k=1}^K(P_{ku}^A - \delta_{ku}) X_{k}^{(N)}\dd\uptau + o(\dd\uptau).
\end{align*}
The remaining terms follow similarly; we omit the straightforward but lengthy algebraic details.
\end{proof}
To prepare for our diffusion limit, we must rescale time; from \eqref{eq:Dij} it is clear that to obtain a nontrivial limit we should let $t = N^{1-\beta}\uptau$. Now introduce the conditional mean vector ${\bfmath{w}}^{(N)}$ and conditional covariance matrix ${\bfmath{s}}^{(N)}$ on this timescale, defined by
\begin{multline}
\lim_{\dd t \to 0}(\dd t)^{-1}\mathbb{E}[\Delta\bfmath{M}^{(N)}\mid \bfmath{M}^{(N)}(t)={\bfmath{m}}] = \\N^{\beta-1}\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\mathbb{E}[\Delta\bfmath{M}^{(N)}\mid \bfmath{M}^{(N)}(\uptau)={\bfmath{m}}]= :{\bfmath{w}}^{(N)}({\bfmath{m}}),\label{eq:w}
\end{multline}
\begin{multline}
\lim_{\dd t \to 0}(\dd t)^{-1}\bbC\textrm{ov}[\Delta\bfmath{M}^{(N)}\mid \bfmath{M}(t)={\bfmath{m}}] =\\ N^{\beta-1}\lim_{\dd\uptau \to 0}(\dd\uptau)^{-1}\bbC\textrm{ov}[\Delta\bfmath{M}^{(N)}\mid \bfmath{M}^{(N)}(\uptau)={\bfmath{m}}] =: N^{\beta-1}{\bfmath{s}}^{(N)}({\bfmath{m}}), \label{eq:s}
\end{multline}
with entries determined by \propref{prop:WF}. Thus, with ${\bfmath{m}} = (x_1,\dots,x_K, \allowbreak y_1,\dots,y_L, \allowbreak d_{11},\dots,d_{KL})$, equations \eref{eq:Xi}--\eref{eq:Dij} show that
\begin{align}
&&{\bfmath{w}}^{(N)}({\bfmath{m}}) &= {\bfmath{w}}({\bfmath{m}}) + O(N^{\beta-1}),\notag\\ \text{where }&&{\bfmath{w}}({\bfmath{m}}) &= \Big(\underbrace{\vphantom{\frac{\rho_\beta}{2}}0,\ldots 0}_{K}, \underbrace{\vphantom{\frac{\rho_\beta}{2}}0, \ldots 0}_{L}, \underbrace{-\frac{\rho_\beta}{2}d_{11},\ldots, -\frac{\rho_\beta}{2}d_{KL}}_{K \times L}\Big)', \label{eq:w2}
\end{align}
with ${\bfmath{s}}^{(N)}({\bfmath{m}}) = {\bfmath{s}}({\bfmath{m}}) + O(N^{-\beta})$ determined in a similar fashion:
\[
{\bfmath{s}}({\bfmath{m}}) = \begin{bmatrix}
{\bfmath{s}}_{\bf XX}({\bfmath{m}}) & {\bfmath{s}}_{\bf XY}({\bfmath{m}}) & {\bfmath{s}}_{\bf XD}({\bfmath{m}})\\
{\bfmath{s}}_{\bf XY}({\bfmath{m}}) & {\bfmath{s}}_{\bf YY}({\bfmath{m}}) & {\bfmath{s}}_{\bf YD}({\bfmath{m}})\\
{\bfmath{s}}_{\bf XD}({\bfmath{m}}) & {\bfmath{s}}_{\bf YD}({\bfmath{m}}) & {\bfmath{s}}_{\bf DD}({\bfmath{m}})
\end{bmatrix},
\]
where
\begin{align*}
[{\bfmath{s}}_{\bf XX}({\bfmath{m}})]_{ik} = {}& x_i(\delta_{ik} - x_k),\\
[{\bfmath{s}}_{\bf YY}({\bfmath{m}})]_{jl} = {}& y_j(\delta_{jl} - y_l),\\
[{\bfmath{s}}_{\bf XY}({\bfmath{m}})]_{ij} = {}& d_{ij},\\
[{\bfmath{s}}_{\bf XD}({\bfmath{m}})]_{i,kl} = {}& d_{kl}(\delta_{ik} - x_i) - x_kd_{il},\\
[{\bfmath{s}}_{\bf YD}({\bfmath{m}})]_{j,kl} = {}& d_{kl}(\delta_{jl} - y_j) - y_ld_{kj},\\
[{\bfmath{s}}_{\bf DD}({\bfmath{m}})]_{ij,kl} = {}& x_iy_j(\delta_{ik} - x_k)(\delta_{jl} - y_l)
{}+ d_{kj}x_iy_l + d_{il}x_ky_j \\
& {}+ d_{ij}(x_ky_l - \delta_{ik}y_l - \delta_{jl}x_k)
{}+ d_{kl}(x_iy_j - \delta_{ik}y_j - \delta_{jl}x_i) \\
&{}+ d_{ij}(\delta_{ik}\delta_{jl} - d_{kl}).
\end{align*}
Notice in particular the different leading orders of the two quantities in \eqref{eq:w} and \eqref{eq:s}:
the mean increments are of $O(1)$ on this timescale while the covariances are of $O(N^{\beta-1})$. It is this difference, which is a consequence of our assumption that the recombination probability $r$ is $O(N^{-\beta})$ for $\beta < 1$, that leads to a novel diffusion limit. Under the usual choice of $\beta = 1$ it is well known that we see convergence to a diffusion process after a linear rescaling of time.
In the special case of a Wright-Fisher model and $K = L = 2$, the diffusion limit for $\bfmath{M}^{(N)}(\lfloor N \uptau \rfloor)$ as $N\to\infty$ was obtained by \citet{oht:kim:1969:GRC, oht:kim:1969:G}. Our interest is however in $\beta \in (0,1)$, for which
$r$ is larger, and the loss of linkage disequilibrium (LD) is subsequently much faster. Intuitively, we should expect such loss to resemble the exponential decay predicted in an infinitely large population, but with small fluctuations about this deterministic behaviour. The diffusion process we define below quantifies these fluctuations precisely.
\subsection{Gaussian diffusion limit of fluctuations in linkage disequilibrium}
We first provide a heuristic description of the diffusion limit. First, observe from \eqref{eq:w} and \eqref{eq:s} that, provided $\bfmath{M}^{(N)}(0) \to \bfmath{M}(0)$ as $N\to\infty$ and that $\beta \in [0,1)$, then
\begin{equation}
\label{eq:M}
\bfmath{M}^{(N)} \overset{d}{\to} \bfmath{M} := \left\{ (\bfmath{X}(0), \bfmath{Y}(0), \bfmath{D}(0)e^{-\rho_\beta t/2})' : t\geq 0\right\}, \qquad N\to\infty,
\end{equation}
the deterministic exponential decay in LD typical of an infinitely large population. See \citet{baa:her:2008} for a formal statement of this law-of-large-numbers type result for the Moran model with recombination. For the corresponding central limit theorem, we seek a diffusion limit for
\begin{equation}
\label{eq:mainrescaling}
\bfmath{U}^{(N)}(t) := r_N[\bfmath{M}^{(N)}(t) - \bfmath{M}(t)],
\end{equation}
for some rescaling $r_N \to \infty$. In our application the appropriate choice is
\[
r_N := N^{(1-\beta)/2},
\]
which can be regarded as the one on which both recombination and genetic drift are observable on the fastest timescale \citep{jen:son:2012:AAP}. We will assume this scaling henceforward. To find the limit $\bfmath{U} = \lim_{N\to\infty} \bfmath{U}^{(N)}$, write
\begin{multline}
\label{eq:maindecomposition}
\bfmath{U}^{(N)}(t) = r_N\Bigg[[\bfmath{M}^{(N)}(0) - \bfmath{M}(0)] \\
{}+ \int_0^t [{\bfmath{w}}^{(N)}(\bfmath{M}^{(N)}(s)) - {\bfmath{w}}(\bfmath{M}(s))]\dd s + \Ma^{(N)}(t)\Bigg],
\end{multline}
where
\[
\Ma^{(N)}(t) := \bfmath{M}^{(N)}(t) - \bfmath{M}^{(N)}(0) - \int_0^t {\bfmath{w}}^{(N)}(\bfmath{M}^{(N)}(s)) \dd s
\]
describes the deviations of $\bfmath{M}^{(N)}(t)$ from its expected behaviour and is a martingale. It suffices to characterize the limits of each of the three grouped terms on the right of \eqref{eq:maindecomposition}. For the first term we assume that it converges to a limit, $\bfmath{U}^{(N)}(0) \overset{d}{\to} \bfmath{U}(0)$ as $N \to\infty$. For the second term, from \eqref{eq:w2} we should expect
\begin{align}
\MoveEqLeft r_N\int_0^t [{\bfmath{w}}^{(N)}(\bfmath{M}^{(N)}(s)) - {\bfmath{w}}(\bfmath{M}(s))]\dd s \notag\\
&= r_N\int_0^t\left[-\frac{\rho_\beta}{2}[\bfmath{M}^{(N)}(s) - \bfmath{M}(s)] +O(N^{\beta-1})\right] \dd s\notag\\
&= \int_0^t \left[-\frac{\rho_\beta}{2}\bfmath{U}^{(N)}(s) +O(N^{(\beta-1)/2})\right]\dd s \notag\\
&\overset{d}{\to} -\frac{\rho_\beta}{2}\int_0^t \bfmath{U}(s) \dd s, \qquad N\to\infty. \label{eq:middleterm}
\end{align}
Finally, we obtain a complete description of the limit $r_N\Ma^{(N)} \overset{d}{\to} \Ma$ as $N\to\infty$ by an application of the martingale central limit theorem \citep[Theorem 7.1.4]{eth:kur:1986}; we find
\[
\Ma(t) = \int_0^t \bfmath{\gs}(\bfmath{M}(s)) \dd\bfmath{W}(s),
\]
where $\bfmath{\gs}\bfgs' = {\bfmath{s}}$, and $\bfmath{W}$ is a $(KL-1)$-dimensional Brownian motion. In summary then, we expect $\bfmath{U}$ to satisfy
\begin{equation}
\label{eq:U}
\bfmath{U}(t) = \bfmath{U}(0) - \frac{\rho_\beta}{2}\int_0^t \bfmath{U}(s) \dd s + \int_0^t \bfmath{\gs}(\bfmath{M}(s)) \dd\bfmath{W}(s).
\end{equation}
Our main result formalizes this argument, as follows.
\begin{theorem}
\label{thm:nor:1975}
Suppose that $\bfmath{U}^{(N)}(0) \overset{d}{\to} \bfmath{U}(0)$ as $N\to\infty$. Then for each $t >0$, as $N\to\infty$,
\[
\sup_{s\leq t}|\bfmath{M}^{(N)}(s) - \bfmath{M}(s)| \overset{d}{\to} 0;
\]
$N^{(1-\beta)/2}\Ma^{(N)} \overset{d}{\to} \Ma$, where $\Ma$ has Gaussian, independent increments with mean zero, and with
\begin{equation}
\label{eq:martingale-covariance}
\mathbb{E}[\Ma(t)\Ma(t)'] = \int_0^t {\bfmath{s}}(\bfmath{M}(s)) \dd s;
\end{equation}
and $\bfmath{U}^{(N)} \overset{d}{\to} \bfmath{U}$, satisfying \eqref{eq:U}.
\end{theorem}
\begin{proof}[Proof of \thmref{thm:nor:1975}]
This is an application of a central limit theorem for density dependent population processes; for textbook coverage see \citet[Chapter 11]{eth:kur:1986} and for a recent treatment see \citet{kan:etal:2014}. We apply Theorem 2.11 of \citet{kan:etal:2014}. To do so we need to validate each of the assertions that led to \eqref{eq:U} above by checking the following sufficient conditions (i)--(iv).
(\citet[][Theorem 2.11]{kan:etal:2014} is rather more general than is required here: it permits the state space of $\bfmath{M}^{(N)}$ to be unbounded, and for $\bfmath{M}^{(N)}$ to depend on other processes that evolve on faster timescales than that of the diffusion. We omit those conditions which are not needed.)
\noindent \emph{(i) {\bf The Moran process converges to an identifiable, deterministic limit.} This is guaranteed by the following:
the infinitesimal generator ${\mathcal A}_N$ of $\bfmath{M}^{(N)}$ satisfies
\[
\lim_{N\to\infty} \sup_{{\bfmath{m}} \in
\Delta_{KL-1}^{(N)}}\left|{\mathcal A}_N f({\bfmath{m}}) - {\mathcal A} f({\bfmath{m}})\right| = 0, \qquad f \in {\mathcal D}({\mathcal A}),
\]
for a generator ${\mathcal A}$ with domain ${\mathcal D}({\mathcal A})$.}
\noindent \emph{(ii) {\bf Fluctuations about the deterministic limit are well behaved.} More precisely, $\bfmath{R}^{(N)}$ is a local martingale and the covariations processes $[\bfmath{M}^{(N)}]_t \overset{d}{\to} 0$.}
\noindent \emph{(iii) {\bf Contributions of $O(r_N^{-1})$ to the error ${\bfmath{w}}^{(N)} - {\bfmath{w}}$ can be identified.} These would contribute to the limiting drift of $\bfmath{U}(t)$, and a sufficient condition to identify them is: there exists a continuous function $\bfmath{G}_0:\Delta_{KL-1} \to \mathbb{R}^{\Lambda}$ (recall $\Lambda = K+L+KL$) such that
\[
\lim_{N\to\infty} \sup_{{\bfmath{m}}\in
\Delta_{KL-1}^{(N)}}\left|r_N\left[{\bfmath{w}}^{(N)}({\bfmath{m}}) - {\bfmath{w}}({\bfmath{m}})\right] - \bfmath{G}_0({\bfmath{m}})\right| = 0.
\]}
\noindent \emph{(iv) {\bf The martingale central limit theorem applies to $r_N \bfmath{R}^{(N)}$.} This is guaranteed by the following:
\begin{equation}
\label{eq:jumpsize}
\lim_{N\to\infty}\mathbb{E}\left[\sup_{s\leq t} r_N \left| \bfmath{M}^{(N)}(s) - \bfmath{M}^{(N)}(s-) \right|\right] = 0,
\end{equation}
and there exists a continuous $\bfmath{G}:\Delta_{KL-1} \to \mathbb{R}^{\Lambda\times \Lambda}$ such that for each $t > 0$,
\begin{equation}
\label{eq:covariation}
r_N^2[\bfmath{M}^{(N)}]_t - \int_0^t \bfmath{G}(\bfmath{M}^{(N)}(s)) \dd s \overset{d}{\to} 0.
\end{equation}
}We address each of these requirements in turn.
\emph{(i)} Convergence of ${\mathcal A}_Nf({\bfmath{m}})$ to ${\mathcal A} f({\bfmath{m}}) := {\bfmath{w}}.\nabla f({\bfmath{m}})$, the generator of $\bfmath{M}$ [see \eqref{eq:M}], is immediate from \propref{prop:WF}. Convergence is uniform in ${\bfmath{m}}$ because the $O(N^{-\beta})$ terms in \propref{prop:WF} have coefficients that are polynomials in $\bfmath{M}^{(N)}$ on a compact space.
\emph{(ii)} Since the state space is bounded, for $\bfmath{R}^{(N)}$ to be a martingale it suffices that the jump rate is uniformly bounded \citep[Proposition 2.1]{kur:1971}, as is the case for the Moran process. The covariations process $[\bfmath{M}^{(N)}]_t \overset{d}{\to} 0$ as a consequence of \eqref{eq:covariation}, verified below.
\emph{(iii)} From \eqref{eq:w2}, $r_N[{\bfmath{w}}^{(N)}({\bfmath{m}}) - {\bfmath{w}}({\bfmath{m}})] = O(N^{(\beta - 1)/2})$, again uniformly in ${\bfmath{m}} \in \Delta_{KL-1}^{(N)}$, so here the appropriate choice is $\bfmath{G}_0 \equiv \bfmath{0}$. Thus, the only relevant contribution to the limit \eqref{eq:middleterm} is from the error ${\bfmath{w}}(\bfmath{M}^{(N)}(s)) - {\bfmath{w}}(\bfmath{M}(s))$ rather than from ${\bfmath{w}}^{(N)}(\bfmath{M}^{(N)}(s)) - {\bfmath{w}}(\bfmath{M}^{(N)}(s))$.
\emph{(iv)} Jumps of any component of $\bfmath{M}^{(N)}$ are bounded in magnitude by $2/N$, so
\[
\sup_{s\leq t} r_N \left| \bfmath{M}^{(N)}(s) - \bfmath{M}^{(N)}(s-) \right| \leq N^{(1-\beta)/2}\cdot \frac{2\Lambda^{1/2}}{N} \to 0, \qquad N\to\infty,
\]
and \eqref{eq:jumpsize} holds. To identify the asymptotic behaviour of $r_N^2[\bfmath{M}^{(N)}]_t$, let
\[
{\mathcal N}^{(N)}_{\bfmath{m}}(t) = {\mathcal Y}_{\bfmath{m}}\left(\int_0^t \lambda^{(N)}_{\bfmath{m}}(\bfmath{M}^{(N)}(s)) \dd s\right)
\]
denote the total number of jumps of the Moran process into state ${\bfmath{m}} \in \Delta_{KL-1}^{(N)}$ by time $t$, where $({\mathcal Y}_{\bfmath{m}} : {\bfmath{m}} \in \Delta_{KL-1}^{(N)})$ is a collection of independent Poisson processes of unit rate and $\lambda_{\bfmath{m}}^{(N)}(\bfmath{M}^{(N)}(s))$ denotes the rate of transition of the process from current state $\bfmath{M}^{(N)}(s)$ to ${\bfmath{m}}$. Then
\begin{align*}
\MoveEqLeft r_N^2[\bfmath{M}^{(N)}]_t = N^{1-\beta}\int_0^t\sum_{{\bfmath{m}} \in \Delta_{KL-1}^{(N)}} [\Delta\bfmath{M}^{(N)}(s)][\Delta\bfmath{M}^{(N)}(s)]' \dd{\mathcal N}^{(N)}_{\bfmath{m}}(s),\\
&\sim N^{1-\beta}\int_0^t\sum_{{\bfmath{m}} \in \Delta_{KL-1}^{(N)}} [{\bfmath{m}}-\bfmath{M}^{(N)}(s)][{\bfmath{m}} -\bfmath{M}^{(N)}(s)]' \lambda_{\bfmath{m}}^{(N)}(\bfmath{M}^{(N)}(s)) \dd s,\\
&\sim \int_0^t {\bfmath{s}}^{(N)}(\bfmath{M}^{(N)}(s)) \dd s,
\end{align*}
by \eqref{eq:s}. Thus we may take $\bfmath{G} = {\bfmath{s}}$ in \eqref{eq:covariation} [$\bfmath{G}$ identifies the moments appearing in \eqref{eq:martingale-covariance}].
\end{proof}
\begin{remark}
One could obtain the same diffusion limit starting from a Wright-Fisher model rather than a Moran model, since the means and covariances of its increments are identical to leading order, up to a rescaling of time. This alternative approach is in some respects less appealing since the Wright-Fisher model, when expressed in continuous time, is non-Markovian. The additional complications raised by this approach have been addressed by \citet{nor:1975:SIAM} \citep[see also][]{eth:nag:1980, eth:nag:1988}, and we have checked that the conditions of his theorems still apply when we introduce recombination to the Wright-Fisher model. The theory of \citet{nor:1975:SIAM} has been used to study strong mutation and selection \citep{nor:1972, nor:1975:SIAM, kap:etal:1988, nag:1986, nag:1990, wak:sar:2009}, and a Gaussian diffusion approximation of a Moran model with strong selection is developed by \citet{fed:etal:2014}, but to the best of our knowledge this is the first time a central limit theorem has been obtained for strong recombination.
\end{remark}
\begin{remark}
The exponential decay of linkage disequilibrium implied by $\bfmath{M}$ [equation \eqref{eq:M}] is a classical result; the above theorem further quantifies the fluctuations about this deterministic behaviour in a fully time-dependent manner. In particular, the definition of $\bfmath{U}$ [equation \eqref{eq:mainrescaling}] shows that fluctuations are of order $N^{(1-\beta)/2}$ on a timescale of $N^{\beta-1}$ units of the Moran process. If we designate the expected lifetime of an individual, $2/N$, as one \emph{generation}, then these fluctuations can be said to occur on a timescale of order $N^\beta$ generations. (This definition of ``generation'' is consistent with $u_A$, $u_B$, and $r$ in \sref{sec:notation} provided we replace $N$ with the effective population size of the Moran model, $N/2$, in the definitions of $\theta_\A$, $\theta_\B$, and $\rho$ \citep[p121]{ewe:2004:I}.)
\end{remark}
\subsection{Stationary distribution}
Although $\bfmath{U}$ is described completely by \eqref{eq:U}, the volatility term $\bfmath{\gs}(\bfmath{M}(t))$ is neither simple nor time-independent. On the other hand, our main interest is in stationary behaviour, and $\bfmath{\gs}(\bfmath{M}(\infty))$ takes on a much simpler form. First note that the components of $\bfmath{U}(t)$ corresponding to each $X_i$ and $Y_j$ undergo Brownian motions (with nonunit volatility), so we restrict our attention to the stationary distribution of the component corresponding to $\bfmath{D}$, which we denote $\bfmath{U}_{\bfmath{D}}$. Conditions of \citet{nor:1975:AAP}
confirm convergence of $\bfmath{U}_{\bfmath{D}}(t)$ to its stationary distribution. Setting $\bfmath{\gs}(\bfmath{M}(s)) = \bfmath{\gs}(\bfmath{M}(\infty))$ in \eqref{eq:U}, we find
\begin{equation}
\label{eq:OU}
\dd\bfmath{U}_{\bfmath{D}} = -\frac{\rho_\beta}{2}\bfmath{U}_{\bfmath{D}}\dd t + \bfmath{\gs}_{\infty} \dd\bfmath{W}(s),
\end{equation}
where $\bfmath{\gs}_{\infty}$ is a \emph{constant} defined by
\[
\bfmath{\gs}_{\infty}\bfmath{\gs}_{\infty}' = {\bfmath{s}}_\infty := {\bfmath{s}}_{\bfmath{D}\bfD}(\bfmath{M}(\infty)) = [X_i(0)Y_j(0)(\delta_{ik}-X_k(0))(\delta_{jl}-Y_l(0))]_{ij,kl}.
\]
The process \eqref{eq:OU} is much simpler to describe. Marginally, $\bfmath{U}_{D_{ij}}$ is an Ornstein-Uhlen\-beck process with damping towards linkage equilibrium at rate $\rho_\beta/2$ and constant volatility $[\bfmath{\gs}_\infty]_{ij,ij}$.
$\bfmath{U}_{\bfmath{D}}$ has stationary distribution
\[
\text{Normal}\left(\bfmath{0}_{KL\times 1}, \frac{{\bfmath{s}}_\infty}{\rho_\beta}\right).
\]
This is a slightly different idea of stationarity than usual, since it depends on $\bfmath{X}(0)$ and $\bfmath{Y}(0)$. An immediate question is: what should be the distributions for $\bfmath{X}(0)$ and $\bfmath{Y}(0)$? We address this by reconsidering the usual two-locus \emph{Wright-Fisher} diffusion limit operating on a slower timescale. We can exploit \eqref{eq:OU} to obtain a simple approximation of this diffusion limit, as follows. First, we have \emph{derived} the Gaussian diffusion approximation
\[
\bfmath{D}(0)e^{-\rho_\beta t/2} + N^{(\beta-1)/2}\bfmath{U}_{\bfmath{D}}(t)
\]
for $\bfmath{D}^{(N)}(t)$.
Thus the stationary distribution of this approximation is
\begin{equation}
\label{eq:OUstationarity}
\text{Normal}\left(\bfmath{0}_{KL\times 1}, \frac{{\bfmath{s}}_\infty}{\rho}\right).
\end{equation}
Notice that this description does not depend on the particular choice of $\beta$. Under the usual ``Wright-Fisher'' regime we treat $\rho$ as fixed.
It remains to specify the stationary distributions for the marginal allele frequencies $\bfmath{X}$ and $\bfmath{Y}$, which we suppose to have reached their usual (independent) stationary distributions in the Wright-Fisher diffusion limit, which we refer to as $\pi_{A}$ and $\pi_{B}$, respectively (and whose respective sampling distributions are $q^A$ and $q^B$). Then we can complete the picture for \eref{eq:OUstationarity} by specifying $(\bfmath{X}(0),\bfmath{Y}(0)) \sim \pi_{A}\otimes \pi_{B}$.
The distribution \eref{eq:OUstationarity} therefore provides a simple, explicit method for the approximate simulation of haplotype frequencies under a stationary, two-locus Wright-Fisher diffusion, which we summarize in the algorithm below. (When mutation is parent independent, as in Remark \ref{rem:jen:son:2009}, $\pi_{A}$ and $\pi_{B}$ take on a particularly simple form, but we note that these distributions are not known in general.)
\\
\fbox{
\begin{minipage}[c]{0.9\textwidth}
{\bf Algorithm to simulate from a Gaussian approximation to the stationary Wright-Fisher diffusion with recombination.}
\begin{enumerate}
\item Simulate marginal allele frequencies at locus A, $\bfmath{X}(0) \sim \pi_{A}$.
\item Independently simulate marginal allele frequencies at locus B,\\ $\bfmath{Y}(0) \sim \pi_{B}$.
\item Conditionally simulate $\bfmath{D}$ from \eref{eq:OUstationarity} given $\bfmath{X}(0)$ and $\bfmath{Y}(0)$.
\item Calculate two-locus haplotype frequencies via
\[
X_{ij} = D_{ij} + X_i(0)Y_j(0), \quad \text{for each }i\in[K], j\in[L].
\]
\end{enumerate}
\end{minipage}}
\\
\subsection{Sampling distribution} The significance of the Gaussian diffusion approximation $\bfmath{U}_{\bfmath{D}}$ is further evident from the following theorem. First we need some further notation. Let
\[
\partition{m} = \left\{{\bfmath{r}}\in \mathbb{N}^{K\times L}: \sum_{i=1}^K\sum_{j=1}^L r_{ij} = m\right\},
\]
for $m \in \mathbb{N}$, and let $\haplist{{\bfmath{r}}} \in ([K]\times [L])^m$ denote a sequence of $m$ haplotypes (in some arbitrary, fixed order) with multiplicities specified by ${\bfmath{r}} \in \partition{m}$. Further let $\haplistA{{\bfmath{r}}} \in [K]^m$ denote the corresponding list of alleles
obtained by looking at the first entry of each element of $\haplist{{\bfmath{r}}}$, and define $\haplistB{{\bfmath{r}}}$ similarly. For $\lambda \in \mathbb{N}$ denote by $\partitiontwo{2\lambda}$ the set of partitions of $[2\lambda]$ with precisely $\lambda$ blocks of size $2$, and write a representative element as $\bfmath{\gx}_{\bfmath{\mu}\bfmath{\nu}} = \{\{\mu_k,\nu_k\}: k=1,\ldots,\lambda\}\in \partitiontwo{2\lambda}$; $\bfmath{\mu} = (\mu_k)$ and $\bfmath{\nu} = (\nu_k)$ are sequences of length $\lambda$.
For $J \subseteq [\lambda]$, denote by $\bfmath{\mu}_J$, $\bfmath{\nu}_J$ the subsequences obtained by looking only at the indices in $J$, and denote by $\haplist{{\bfmath{r}}}_{\bfmath{\mu}}$ the subsequence of $\haplist{{\bfmath{r}}}$ obtained by looking only at the indices in $\bfmath{\mu}$. The matrix of multiplicities of $\haplist{{\bfmath{r}}}_{\bfmath{\mu}}$ is denoted by ${\bfmath{r}}^{(\bfmath{\mu})}$, so that ${\bfmath{r}}^{(\bfmath{\mu})} + {\bfmath{r}}^{(\bfmath{\nu})} = {\bfmath{r}}$. For example, if ${\bfmath{r}} = [\begin{smallmatrix} 1 & 2\\ 0 & 1 \end{smallmatrix}]$ then a representative list of haplotypes is $\haplist{{\bfmath{r}}} = ((1,1),(1,2),(1,2),(2,2))$ with marginal allele lists $\haplistA{{\bfmath{r}}} = (1,1,1,2)$ and $\haplistB{{\bfmath{r}}} = (1,2,2,2)$. Here, $m = 2\lambda = 4$, and $\partitiontwo{4} = \big\{\{\{1,2\},\{3,4\}\}, \{\{1,3\},\{2,4\}\}, \{\{1,4\},\{2,3\}\}\big\}$. Then for example the first element in $\partitiontwo{4}$ is the partition $\bfmath{\gx}_{\bfmath{\mu}\bfmath{\nu}}$ constructed from $\bfmath{\mu} = (1,3)$ and $\bfmath{\nu} = (2,4)$, and so $\haplist{{\bfmath{r}}}_{\bfmath{\mu}} = ((1,1),(1,2))$ and $\haplist{{\bfmath{r}}}_{\bfmath{\nu}} = ((1,2),(2,2))$.
\begin{theorem}
\label{thm:gaussiandiffusion}
Suppose that $\bfmath{X} \sim \pi_{A}$, $\bfmath{Y} \sim \pi_{B}$ independently, and conditional on $\bfmath{X}$ and $\bfmath{Y}$, $\bfmath{D}$ is distributed according to the Gaussian distribution in \eqref{eq:OUstationarity}. Then the sampling distribution is given \emph{exactly} by
{\allowdisplaybreaks
\begin{align}
q_{G}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = {}& \sum_{\lambda=0}^{\lfloor c/2 \rfloor} \frac{1}{\rho^{\lambda}}\sum_{{\bfmath{r}}\in \partition{2\lambda}} \sum_{\bfmath{\gx}\in \partitiontwo{2\lambda}}\left[\prod_{i=1}^K \prod_{j=1}^L\binom{c_{ij}}{r_{ij}}\right]\notag\\
{}& \times \Bigg[\sum_{I \subseteq [\lambda]:\,\haplistA{{\bfmath{r}}}_{\bfmath{\mu}_I} = \haplistA{{\bfmath{r}}}_{\bfmath{\nu}_I}} (-1)^{|I^\complement|}q^A({\bfmath{a}} + {\bfmath{c}}_A-{\bfmath{r}}_A^{(\bfmath{\nu}_I)})\Bigg]\notag\\
& {}\times \Bigg[\sum_{J\subseteq [\lambda]:\,\haplistB{{\bfmath{r}}}_{\bfmath{\mu}_J}=\haplistB{{\bfmath{r}}}_{\bfmath{\nu}_J}}(-1)^{|J^\complement|}q^B({\bfmath{b}} + {\bfmath{c}}_B - {\bfmath{r}}_B^{(\bfmath{\nu}_J)})\Bigg], \label{eq:diffusionexact}\\
= {}& q_0({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) + \frac{q_1({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})}{\rho} + O\left(\frac{1}{\rho^2}\right), \notag
\end{align}}
with $q_0$ and $q_1$ given by \eref{eq:zerothorder} and \eref{eq:firstorder} respectively (and we impose the convention that the empty summations for $\lambda = 0$ have a single term, with $(-1)^{|\varnothing\setminus\varnothing|} = 1$).
\end{theorem}
\begin{proof}
With respect to the diffusion in the transformed co-ordinate system, the sampling distribution is
{\allowdisplaybreaks
\begin{align*}
q_{G}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})
={} & \mathbb{E}\left[\Bigg(\prod_{i=1}^K X_{i}^{a_i}\Bigg)\Bigg(\prod_{j=1}^L Y_{j}^{b_j}\Bigg)\Bigg(\prod_{i=1}^K \prod_{j=1}^L \left[D_{ij} + X_{i}Y_{j}\right]^{c_{ij}}\Bigg)\right], \\
={} & \sum_{m=0}^c \sum_{{\bfmath{r}}\in \partition{m}} \left[\prod_{i=1}^K \prod_{j=1}^L\binom{c_{ij}}{r_{ij}}\right]
\mathbb{E}\left[\Bigg(\prod_{i=1}^K X_{i}^{a_i+c_{i\cdot}-r_{i\cdot}}\Bigg)\right.\\
& \specialcell{\hfill {}\times \left. \Bigg(\prod_{j=1}^L Y_{j}^{b_j+c_{\cdot j} - r_{\cdot j}}\Bigg) \mathbb{E}\left[\prod_{i=1}^K \prod_{j=1}^L D_{ij}^{r_{ij}}\mid \bfmath{X}, \bfmath{Y}\right]\right],}\\
={} & \sum_{\lambda=0}^{\lfloor c/2\rfloor} \sum_{{\bfmath{r}}\in \partition{2\lambda}} \sum_{\bfmath{\gx}\in \partitiontwo{2\lambda}}\left[\prod_{i=1}^K \prod_{j=1}^L\binom{c_{ij}}{r_{ij}}\right]\mathbb{E}\left[\Bigg(\prod_{i=1}^K X_{i}^{a_i+c_{i\cdot}-r_{i\cdot}}\Bigg)\right.\\
& \specialcell{\hfill {}\times \left. \Bigg(\prod_{j=1}^L Y_{j}^{b_j+c_{\cdot j} - r_{\cdot j}}\Bigg) \prod_{k=1}^{\lambda} \mathbb{E}[D_{\haplist{{\bfmath{r}}}_{\mu_k}}D_{\haplist{{\bfmath{r}}}_{\nu_k}}\mid \bfmath{X}, \bfmath{Y}]\right],}\\
={} & \sum_{\lambda=0}^{\lfloor c/2 \rfloor} \frac{1}{\rho^{\lambda}}\sum_{{\bfmath{r}}\in \partition{2\lambda}} \sum_{\bfmath{\gx}\in \partitiontwo{2\lambda}}\left[\prod_{i=1}^K \prod_{j=1}^L\binom{c_{ij}}{r_{ij}}\right] \\
& \times \mathbb{E}\left[\Bigg(\prod_{i=1}^K X_{i}^{a_i+c_{i\cdot}-r_{i\cdot}}\Bigg)\Bigg(\prod_{j=1}^L Y_{j}^{b_j+c_{\cdot j} - r_{\cdot j}}\Bigg)\right.\\
& \specialcell{\hfill \left. {}\times \prod_{k=1}^{\lambda}X_{\haplistA{{\bfmath{r}}}_{\mu_k}}Y_{\haplistB{{\bfmath{r}}}_{\mu_k}}(\delta_{\haplistA{{\bfmath{r}}}_{\mu_k} \haplistA{{\bfmath{r}}}_{\nu_k}} - X_{\haplistA{{\bfmath{r}}}_{\nu_k}})(\delta_{\haplistB{{\bfmath{r}}}_{\mu_k}\haplistB{{\bfmath{r}}}_{\nu_k}} - Y_{\haplistB{{\bfmath{r}}}_{\nu_k}})\right],}\\
={}& \sum_{\lambda=0}^{\lfloor c/2 \rfloor} \frac{1}{\rho^{\lambda}}\sum_{{\bfmath{r}}\in \partition{2\lambda}} \sum_{\bfmath{\gx}\in \partitiontwo{2\lambda}}\left[\prod_{i=1}^K \prod_{j=1}^L\binom{c_{ij}}{r_{ij}}\right] \\
& {}\times \sum_{I \subseteq [\lambda]} (-1)^{|I^\complement|}\delta_{\haplistA{{\bfmath{r}}}_{\bfmath{\mu}_I}\haplistA{{\bfmath{r}}}_{\bfmath{\nu}_I}}\sum_{J\subseteq [\lambda]}(-1)^{|J^\complement|}\delta_{\haplistB{{\bfmath{r}}}_{\bfmath{\mu}_J}\haplistB{{\bfmath{r}}}_{\bfmath{\nu}_J}}\\
& {} \times \mathbb{E}\left[\Bigg(\prod_{i=1}^K X_{i}^{a_i+c_{i\cdot}-r^{(\bfmath{\nu}_I)}_{i\cdot}}\Bigg)\Bigg(\prod_{j=1}^L Y_{j}^{b_j+c_{\cdot j} - r^{(\bfmath{\nu}_J)}_{\cdot j}}\Bigg) \right],
\end{align*}}
The second equality follows from the multinomial theorem and the tower property, the third equality follows from Isserlis' theorem \citep{mic:etal:2011},
and the fourth equality follows from \eref{eq:OUstationarity}:
\[
\mathbb{E}[D_{ij}D_{kl}\mid \bfmath{X}, \bfmath{Y}] = \frac{1}{\rho}X_iY_j(\delta_{ik} - X_k)(\delta_{jl} - Y_l).
\]
The fifth equality follows from expanding the final product (using the convention $\delta_{\varnothing\varnothing} = 1$), while \eref{eq:diffusionexact} follows from $(\bfmath{X},\bfmath{Y}) \sim \pi_A \otimes \pi_B$. The equalities still hold for $\lambda = 0$ provided we take $\prod_\varnothing = 1$.
Extracting the two leading order terms $\lambda=0$ and $\lambda = 1$, the expression simplifies to
\begin{align*}
q_{G}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = {}& \mathbb{E}\left[\Bigg(\prod_{i=1}^K X_{i}^{a_i+c_{i\cdot}}\Bigg)\Bigg(\prod_{j=1}^L Y_{j}^{b_j+c_{\cdot j}}\Bigg)\right]\\
& {}+ \frac{1}{\rho}\sum_{k,u=1}^K\sum_{l,v=1}^L \frac{c_{kl}(c_{uv} - \delta_{ku}\delta_{lv})}{2} \mathbb{E}\left[\Bigg(\prod_{i=1}^K X_{i}^{a_i+c_{i\cdot}-\delta_{iu}}\Bigg)\right.\\
& \specialcell{\hfill \left. {}\times \Bigg(\prod_{j=1}^L Y_{j}^{b_j+c_{\cdot j} - \delta_{jv}}\Bigg) (\delta_{ku} - X_{u})(\delta_{lv} - Y_{v})\right] + O\left(\frac{1}{\rho^2}\right),}\\
= {}& q_0({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) + \frac{q_1({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})}{\rho} + O\left(\frac{1}{\rho^2}\right),
\end{align*}
as required.
\end{proof}
\subsection{Accuracy of the diffusion process}
A natural question to ask is: to what extent does the process of \thmref{thm:gaussiandiffusion} capture the dynamics of the full process? To address this we consider the accuracy of the sampling distribution \eref{eq:diffusionexact} as an approximation to the ``true'' distribution, $q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$. For moderate sample sizes it is possible to compute the latter as the solution to a system of recursive equations \citep{gol:1984, eth:gri:1990, jen:son:2009:G}. The number of summands in \eref{eq:diffusionexact} grows rapidly with $\lambda$ (as long as $\lambda \leq \lfloor \frac{c}{2}\rfloor$), so we define an approximate sampling distribution $q_{G}^{(\lambda)}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$ by truncating the outer sum in \eref{eq:diffusionexact} at a fixed index $\lambda$. This is analogous to the asymptotic sampling formulae for the full model which are obtained by truncating equation \eref{eq:main} \citep{jen:son:2012:AAP}. As our measure of accuracy we define the relative error,
\begin{equation}
\label{eq:relativeerror}
\eG{\lambda} = \left| \frac{Q_{G}^{(\lambda)}(\bfmath{0},\bfmath{0},{\bfmath{c}}) - q(\bfmath{0},\bfmath{0},{\bfmath{c}})}{q(\bfmath{0},\bfmath{0},{\bfmath{c}})}\right| \times 100\%,
\end{equation}
where $Q_{G}^{(\lambda)}(\bfmath{0},\bfmath{0},{\bfmath{c}})$ is the staircase Pad\'e approximant to $q_{G}^{(\lambda)}(\bfmath{0},\bfmath{0},{\bfmath{c}})$. \citep[The former is used for its superior convergence properties; see][for details.]{jen:son:2012:AAP} We define $\eT{\lambda}$ analogously, replacing $Q_{G}^{(\lambda)}(\bfmath{0},\bfmath{0},{\bfmath{c}})$ in \eref{eq:relativeerror} with the Pad\'e approximant to the partial sum of \eref{eq:main}, computed up to $O(\rho^{-(\lambda+1)})$ by the method of \citet{jen:son:2012:AAP}.
We computed the distribution of $\eG{\lambda}$ and of $\eT{\lambda}$ across all sample configurations of size $c=20$ for which both alleles are observed at each locus; results are shown in \tref{tab:accuracy}. For a collection of this size it was straightforward to compute up to $\lambda = 6$ for every possible sample configuration. Using a partial sum to approximate \eref{eq:main} contributes to both errors; $\eG{\lambda}$ has additional contributions reflecting its use of an approximate \emph{model}. Of course, the two errors agree up to $\lambda =1$. However, \tref{tab:accuracy} shows that they are comparable more broadly, particularly for large recombination rates. As $\lambda$ increases, $Q_{G}^{(\lambda)}(\bfmath{0},\bfmath{0},{\bfmath{c}})$ converges rapidly
(even without Pad\'e summation; not shown), and becomes a reasonable approximation to $q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$. For example, for $\rho = 50$, $Q^{(6)}_{G}(\bfmath{0},\bfmath{0},{\bfmath{c}})$ is within $10\%$ of $q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$ with probability $0.79$, though it is within $1\%$ only with probability $0.50$. When we consider the highest levels of accuracy, as in $\Phi(1)$ in \tref{tab:accuracy}, $\eG{\lambda}$ actually increases with $\lambda$ when $\lambda > 1$. This suggests that the Gaussian model typically cannot approximate the true model to the same level of precision as a first order asymptotic approximation of the true model, though its behaviour as a coarser approximation (as reflected in the columns for $\Phi(100)$, for example) is comparable.
\begin{table}[t]
\begin{center}
\caption{\label{tab:accuracy}Cumulative distribution $\Phi(x) = \mathbb{P}(\mathring{e}^{(\lambda)} < x\%)$ (where $\mathring{e}^{(\lambda)}$ denotes either $\eG{\lambda}$ or $\eT{\lambda}$ as defined in the main text), for all samples of size $20$ dimorphic at both loci.}
\label{tab:main}
\begin{tabular}{ccccc|cccc}
& & \multicolumn{3}{c}{$\rho = 25$} & \multicolumn{3}{c}{$\rho = 50$}\\
\hline
& Type&&&\\
$\lambda$ & of sum & $\Phi(1)$ & $\Phi(10)$ & $\Phi(100)$& $\Phi(1)$ & $\Phi(10)$ & $\Phi(100)$ \\\hline
0 & True & 0.39 & 0.58 & 1.00 & 0.49 & 0.63 & 1.00 \\
& Gaussian & 0.39 & 0.58 & 1.00 & 0.49 & 0.63 & 1.00 \\
1 & True & 0.51 & 0.75 & 0.96 & 0.59 & 0.84 & 0.99 \\
& Gaussian & 0.51 & 0.75 & 0.96 & 0.59 & 0.84 & 0.99 \\
2 & True & 0.59 & 0.91 & 0.97 & 0.77 & 0.98 & 1.00 \\
& Gaussian & 0.50 & 0.73 & 0.97 & 0.50 & 0.86 & 1.00 \\
4 & True & 0.83 & 0.99 & 1.00 & 0.95 & 1.00 & 1.00 \\
& Gaussian & 0.51 & 0.72 & 1.00 & 0.50 & 0.80 & 1.00 \\
6 & True & 0.89 & 0.99 & 1.00 & 0.99 & 1.00 & 1.00 \\
& Gaussian & 0.49 & 0.71 & 0.99 & 0.50 & 0.79 & 1.00 \\
\hline\\[-5pt]
& & \multicolumn{3}{c}{$\rho = 100$} & \multicolumn{3}{c}{$\rho = 200$}\\
\hline
& Type&&&\\
$\lambda$ & of sum & $\Phi(1)$ & $\Phi(10)$ & $\Phi(100)$& $\Phi(1)$ & $\Phi(10)$ & $\Phi(100)$ \\\hline
0 & True & 0.50 & 0.72 & 1.00 & 0.54 & 0.95 & 1.00 \\
& Gaussian & 0.50 & 0.72 & 1.00 & 0.54 & 0.95 & 1.00 \\
1 & True & 0.74 & 0.95 & 1.00 & 0.90 & 0.99 & 1.00 \\
& Gaussian & 0.74 & 0.95 & 1.00 & 0.90 & 0.99 & 1.00 \\
2 & True & 0.95 & 1.00 & 1.00 & 1.00 & 1.00 & 1.00 \\
& Gaussian & 0.64 & 0.99 & 1.00 & 0.85 & 1.00 & 1.00 \\
4 & True & 1.00 & 1.00 & 1.00 & 1.00 & 1.00 & 1.00 \\
& Gaussian & 0.64 & 0.99 & 1.00 & 0.83 & 1.00 & 1.00 \\
6 & True & 1.00 & 1.00 & 1.00 & 1.00 & 1.00 & 1.00 \\
& Gaussian & 0.64 & 0.99 & 1.00 & 0.83 & 1.00 & 1.00 \\
\hline
\end{tabular}
\end{center}
\end{table}
\section{Coalescent process}
\label{sec:coalescent}
\subsection{A coupling argument}
In this section we derive a coalescent process which is much simpler than the ARG but whose sampling distribution agrees with \eref{eq:zerothorder} and \eref{eq:firstorder}. We first provide an informal description. Let ${\mathcal C}^{(\rho)}_{a,b,c}(t)$ denote the standard, neutral, two-locus coalescent process a time $t$ back from a sample taken at time $t=0$, with $a$, $b$, and $c$ counting the three types of sample as defined in \sref{sec:notation}.
Recombination occurs at the usual rate of $\rho c/2$, where $\rho = 2Nr$. Lineages ancestral to the three types
are sometimes referred to as representing \emph{left half-fragments}, \emph{right half-fragments}, and \emph{full fragments}, respectively. Our strategy is to define a coupling on a joint probability space for the pair of processes $({\mathcal C}^{(\rho)} = ({\mathcal C}^{(\rho)}_{a,b,c}(t):t \geq 0), {\mathcal D}^{(\infty)} = ({\mathcal D}^{(\infty)}_{a,b,c}(t)): t\geq 0))$, where ${\mathcal D}^{(\infty)}$ is a simple process closely related to ${\mathcal C}^{(\infty)}$ and defined below. ${\mathcal C}^{(\rho)}(\omega )$ is said to be coupled to ${\mathcal D}^{(\infty)}(\omega )$ if the two realizations have the same marginal coalescent tree at locus A and the same marginal coalescent tree at locus B. Since it is the marginal trees which govern the mutation process at each locus, coupled processes therefore have the same sampling distribution. (There should be no ambiguity arising from the fact that our coupling is not on pairs of realizations but on pairs of equivalence classes, where an equivalence class of ${\mathcal C}^{(\rho)}$ or of ${\mathcal D}^{(\infty)}$ is a set of realizations with the same marginal tree at locus A and the same marginal tree at locus B.)
A complete description of a coalescent process is one taking values in partitions of $[n]$, as introduced by \citet{kin:1982:SPA}, with natural extensions to incorporate recombination. We opt instead to represent ${\mathcal C}^{(\rho)}$ only by its \emph{ancestral} process; that is, as a birth-death process on the \emph{number} of each type of lineage. Such a process is studied in depth by \citet{eth:gri:1990} and \citet{gri:1991}. In what follows it is understood implicitly that for any given realization of the ancestral process one could reconstruct a complete coalescent process---an ARG---given some additional independent randomness. Provided the ancestral processes of ${\mathcal C}^{(\rho)}$ and ${\mathcal D}^{(\infty)}$ remain coupled, then it is also always possible to couple their respective \emph{coalescent} processes. For example, a decrease by one in the ancestral process corresponds to a coalescence event in the coalescent process, which can be realized by merging two uniformly chosen blocks in the partition of $[n]$. A coupling of two \emph{ancestral} processes lets us couple the corresponding \emph{coalescent} processes if we always pick the same pair of blocks to merge in the two processes. With this kept in mind, it is sufficient for the argument developed below to consider the simpler ancestral process representation.
Recall the two-locus ancestral process for the coalescent with recombination: Going backwards in time, each pair of lineages coalesces independently at rate $1$, and each lineage ancestral at both loci recombines at rate $\rho/2$. When two lineages coalesce, they are replaced with a single lineage, and this lineage is ancestral at a given locus if either of its two progenitors were ancestral at this locus. Thus for example, with $a$, $b$, and $c$ defined as above the total rate of coalescence involving one left-half fragment and one right-half fragment is $ab$, resulting in a transition of the form $(a,b,c) \mapsto (a-1,b-1,c+1)$. The remaining transitions are given in \tref{tab:coupling}. We can now make the following concise definition.
\begin{definition}
The ancestral process ${\mathcal C}^{(\rho)} = ({\mathcal C}^{(\rho)}_{a,b,c}(t):t \geq 0)$ is a continuous-time Markov process on $\mathbb{N}^4$ such that ${\mathcal C}^{(\rho)}_{a,b,c}(0) = (a,b,c,c)$ a.s., and with infinitesimal generator
\begin{align}
\mathscr{L} f(a,b,c,c) = {} & \frac{\rho c}{2}f(a+1,b+1,c-1,c-1) + \binom{c}{2}f(a,b,c-1,c-1) \notag\\
& {}+ R_{a,b,c,c}\mathscr{G} f(a,b,c,c) - \left[\frac{\rho c}{2} + \binom{c}{2} + R_{a,b,c,c}\right]f(a,b,c,c), \label{eq:genrho}
\end{align}
where
\begin{align*}
R_{a,b,c,d} = {}& ab + ac + bd + \binom{a}{2} + \binom{b}{2},\\
\mathscr{G} f(a,b,c,d) = {}& \frac{1}{2R_{a,b,c,d}}[2abf(a-1,b-1,c+1,d+1) \\
& {}+ a(a+2c-1)f(a-1,b,c,d) + b(b+2d-1)f(a,b-1,c,d)],
\end{align*}
and $f: \mathbb{N}^4 \to \mathbb{R}$ is an appropriate test function.
\end{definition}
Regard the third and fourth entries in $f$ as the number of left- and right- halves of full fragments; these entries are always equal. This representation is seemingly redundant, but it will make the coupling with the corresponding process ${\mathcal D}^{(\infty)}$ (for which we allow $c\neq d$) transparent. We will define ${\mathcal D}^{(\infty)}$ via the following recipe. First, take ${\mathcal C}^{(\rho)}$ and let $\rho\to\infty$. Ordinarily, ${\mathcal C}^{(\infty)}_{a,b,c}(0)$ moves instantaneously to the state ${\mathcal C}^{(\infty)}_{a+c,b+c,0}(0+)$ and evolves thereafter according to $\mathscr{L} f(a+c,b+c,0,0)$. However, our second step is to make a notational change: we reuse the third and fourth entries of $f$ by separately tracking the half-fragment lineages that \emph{originated} as full fragments: we write it as a process initiated at $(a,b,c,c)$ and evolving according to the generator
\begin{multline}
\mathscr{L}^{(\infty)} f(a,b,c,d) = \binom{c}{2}f(a,b,c-1,d) + \binom{d}{2}f(a,b,c,d-1)\\
{}+ R_{a,b,c,d}\mathscr{G} f(a,b,c,d) - \left[\binom{c}{2} + \binom{d}{2} + R_{a,b,c,d}\right]f(a,b,c,d).\label{eq:geninf}
\end{multline}
Third, we
introduce an \emph{artificial} recombination process which induces transitions of the form $(a,b,c,c) \mapsto (a+1,b+1,c-1,c-1)$ at rate $\rho c/2$. This does not reflect any concrete evolutionary dynamic but merely acts as a mathematical device to facilitate a coupling between the two processes. (As a minor technical detail, we should like to allow the process ultimately to reach a state of the form $(a,b,0,0)$. We therefore make a minor adjustment, below, to this artificial process to allow for it to act even if one of $c$ or $d$ is $0$.) We therefore have the following definition.
\begin{definition}
The ancestral process ${\mathcal D}^{(\infty)} = ({\mathcal D}^{(\infty)}_{a,b,c}(t):t \geq 0)$ is a continuous-time Markov process on $\mathbb{N}^4$ such that ${\mathcal D}^{(\infty)}_{a,b,c}(0) = (a,b,c,c)$ a.s., and with infinitesimal generator
\begin{multline}
\mathscr{H}^{(\infty)} f(a,b,c,d) := \mathscr{L}^{(\infty)} f(a,b,c,d) \\
{}+ \frac{\rho \max\{c,d\}}{2}[f(a+\mathbb{I}\{c > 0\},b+\mathbb{I}\{d > 0\},c-\mathbb{I}\{c > 0\},d-\mathbb{I}\{d > 0\}) \\{}- f(a,b,c,d)],\label{eq:geninf2}
\end{multline}
where $f: \mathbb{N}^4 \to \mathbb{R}$ is an appropriate test function.
\end{definition}
Transitions of this process are also summarized in \tref{tab:coupling}, and henceforth we will refer to the numberings of each type of transition given in the table. It is important to keep in mind that although $\rho$ appears as a parameter in \eqref{eq:geninf2}, the process ${\mathcal D}^{(\infty)}$ acts as if the two loci are independent. The process with rate depending on $\rho$ is simply an artificial relabelling of lineages. A key observation is that this artificial process does not affect the distribution of the marginal coalescent trees, so ${\mathcal C}^{(\infty)}$ and ${\mathcal D}^{(\infty)}$ have the same sampling distribution.
\begin{table}[t]
\begin{center}
\caption{\label{tab:coupling}Transition rates of events in the two ancestral processes ${\mathcal C}^{(\rho)}$ and ${\mathcal D}^{(\infty)}$.}
\begin{tabular}{cc|cc}
\hline
& Transition & \multicolumn{2}{c}{Rate}\\
Type & $(a,b,c,d) \mapsto $& ${\mathcal C}^{(\rho)}$
& ${\mathcal D}^{(\infty)}$\\
\hline
{\sc I} & $(a,b,c-1,d-1)$ & $c(c-1)/2^*$
& $0$\\
{\sc II} & $(a,b,c-1,d)$ & $0$
& $c(c-1)/2$ \\
{\sc III} & $(a,b,c,d-1)$ & $0$
& $d(d-1)/2$ \\
{\sc IV} & $(a-1,b,c,d)$ & $a(a+2c-1)/2$
& $a(a+2c-1)/2$ \\
{\sc V} & $(a,b-1,c,d)$ & $b(b+2d-1)/2$
& $b(b+2d-1)/2$ \\
{\sc VI} & $(a-1,b-1,c+1,d+1)$ & $ab$
& $ab$\\
{\sc VII} & $(a+\mathbb{I}\{c > 0\},b+\mathbb{I}\{d > 0\},$~~ \\ & ~~$c-\mathbb{I}\{c > 0\},d-\mathbb{I}\{d > 0\})$ & $\rho c/2^*$
& $\rho \max\{c,d\}/2$\\
\hline
\end{tabular}
\end{center}
$^*${\small Defined only when $c = d$.}
\end{table}
To summarize, we have defined two Markov processes on $\mathbb{N}^4$, ${\mathcal C}^{(\rho})$ and ${\mathcal D}^{(\infty)}$,
which describe two-locus ancestral processes going backwards in time and with respective generators $\mathscr{L}$ and $\mathscr{H}^{(\infty)}$. $\mathscr{L}$ is the generator of a standard process with recombination parameter $\rho$. $\mathscr{H}^{(\infty)}$ is the generator of a standard process with recombination parameter $\infty$ and with the additional properties that left half-fragments are recorded in two categories (of multiplicity $a$ and $c$), right half-fragments are recorded in two categories (of multiplicity $b$ and $d$), and there is an artificial movement of pairs from the latter to the former as if they were still full fragments. This somewhat contrived definition has an important advantage: it is a simple matter to attempt to couple the two processes by matching each kind of event in the two generators whenever possible. A recombination event in ${\mathcal C}^{(\rho)}_{a,b,c}(t)$ can be matched by an artificial recombination event in ${\mathcal D}^{(\infty)}_{a,b,c}(t)$, a coalescence of type {\sc IV}{}
in ${\mathcal C}^{(\rho)}_{a,b,c}(t)$ can be matched by a coalescence of type {\sc IV}{} in ${\mathcal D}^{(\infty)}_{a,b,c}(t)$, and so on.
The aforementioned description is a probabilistic coupling, which may or may not succeed since not all events can be paired off in this way. Comparing \eref{eq:genrho} and \eref{eq:geninf2}, we see that a coupling will fail if there is a type {\sc I}{} transition
in ${\mathcal C}^{(\rho)}$ or if there is a type {\sc II}{} or type {\sc III}{} transition
in ${\mathcal D}^{(\infty)}$.
Define the failure times
\begin{align*}
T^{(1)}_{a,b,c} &:= \inf\{t \geq 0: {\mathcal C}^{(\rho)}_{a,b,c}(t) = {\mathcal C}^{(\rho)}_{a,b,c}(t-) - (0,0,1,1)\},\\
T^{(2)}_{a,b,c} &:= \inf\{t \geq 0: {\mathcal D}^{(\infty)}_{a,b,c}(t) = {\mathcal D}^{(\infty)}_{a,b,c}(t-) - (0,0,1,0)\},\\
T^{(3)}_{a,b,c} &:= \inf\{t \geq 0: {\mathcal D}^{(\infty)}_{a,b,c}(t) = {\mathcal D}^{(\infty)}_{a,b,c}(t-) - (0,0,0,1)\},
\end{align*}
and
\begin{multline*}
T^{\text{\tiny MRCA}}_{a,b,c} := \inf\Big\{t \geq 0: {\mathcal C}^{(\rho)}_{a,b,c}(s) = {\mathcal D}^{(\infty)}_{a,b,c}(s) \quad\forall s \leq t, \\ {\mathcal C}^{(\rho)}_{a,b,c}(t) \in \{(1,1,0,0),(0,0,1,1)\}\Big\},
\end{multline*}
the first time that both loci find a most recent common ancestor in the coupled processes (with the convention $\inf \varnothing = \infty$). If $T^{\text{\tiny MRCA}}_{a,b,c} < \min\{T^{(1)}_{a,b,c}$, $T^{(2)}_{a,b,c}$, $T^{(3)}_{a,b,c}\}$, we say that the coupling has been \emph{successful}. We are now in a position to verify the observation made in \sref{sec:intro}: that we need consider whether or not a coupling has been successful only as far back as the first time that no lineages ancestral to both loci survive. For if we reach this point then, even further back in time, jointly ancestral lineages may arise again temporarily (with $c \geq 1$), but the coupling can fail only in the unlikely [i.e.\ $O(\rho^{-2})$] event that $c \geq 2$. We formalize this argument in the following lemma.
\begin{lemma}
\label{lem:c=0}
If $c \in\{0,1\}$, the coupling between ${\mathcal C}^{(\rho)}$ and ${\mathcal D}^{(\infty)}$ fails with probability $O(\rho^{-2})$, as $\rho\to\infty$.
\end{lemma}
\begin{proof}
The three events causing the coupling to fail occur at rates proportional to $\binom{c}{2}$ and thus require $c \geq 2$. For the pair $({\mathcal C}^{(\rho)}_{a,b,1}, {\mathcal D}^{(\infty)}_{a,b,1})$, we therefore first need to see a transition of the form $(a',b',1,1) \mapsto (a'-1,b'-1,2,2)$ for some $a',b'$, followed by one of the transitions causing the coupling to fail. Reading off the rates from the generators, each of these transitions occurs with probability $O(\rho^{-1})$. The case $c=0$ is similar, first needing a transition of the form $(a',b',0,0) \mapsto (a'-1,b'-1,1,1)$ whose probability is of $O(1)$.
\end{proof}
\begin{lemma}
\label{lem:coupling}
The coupling between ${\mathcal C}^{(\rho)}$ and ${\mathcal D}^{(\infty)}$ fails with the following probabilities:
\begin{equation}
\label{eq:failures}
\mathbb{P}(I^{(k)}) = \frac{1}{\rho}\binom{c}{2} + O\left(\frac{1}{\rho^2}\right) \quad \text{as }\rho\to\infty, \quad k = 1,2,3,
\end{equation}
where $I^{(k)} := \{T^{(k)}_{a,b,c} < T^{\text{\tiny MRCA}}_{a,b,c}\}$. Moreover, $\mathbb{P}(I^{(k_1)} \cap I^{(k_2)}) = O(\rho^{-2})$ for $k_1 \neq k_2$.
\end{lemma}
\begin{proof}
For $k=1$, by \lemmaref{lem:c=0} it is enough to show that
\[
\mathbb{P}(T^{(1)}_{a,b,c} < U^{(1)}_{a,b,c}) = \frac{1}{\rho}\binom{c}{2} + O\left(\frac{1}{\rho^2}\right),
\]
where
\begin{align*}
U^{(1)}_{a,b,c} := {}& \inf\left\{t \geq 0:
{\mathcal C}^{(\rho)}_{a,b,c}(t) \in \{(a',b',0,0) : a',b'\in \mathbb{N}\}\right\}
\end{align*}
is the first time ${\mathcal C}^{(\rho)}$ reaches $c=0$. We proceed by induction on $c$; \lemmaref{lem:c=0} provides the base cases $c\in\{0,1\}$. First note that for any $c \geq 1$,
\begin{equation}
\label{eq:T1U1}
\mathbb{P}(T^{(1)}_{a,b,c} < U^{(1)}_{a,b,c}) = O\left(\frac{1}{\rho}\right),
\end{equation}
since this event requires at least one transition that is not a recombination. Reading off the relevant probabilities from \eref{eq:genrho}, we have for $c \geq 2$:
\begin{align*}
\mathbb{P}(T^{(1)}_{a,b,c} < U^{(1)}_{a,b,c}) = {}& \frac{\frac{\rho c}{2}}{\frac{\rho c}{2} + \binom{c}{2} + R_{a,b,c,c}}\cdot\mathbb{P}(T^{(1)}_{a+1,b+1,c-1} < U^{(1)}_{a+1,b+1,c-1}) \\
& {}+ \frac{ab}{\frac{\rho c}{2} + \binom{c}{2} + R_{a,b,c,c}}\cdot \mathbb{P}(T^{(1)}_{a-1,b-1,c+1} < U^{(1)}_{a-1,b-1,c+1})\\
& {}+ \frac{\binom{c}{2}}{\frac{\rho c}{2} + \binom{c}{2} + R_{a,b,c,c}}\cdot 1+ O\left(\frac{1}{\rho^{2}}\right),\\
= {}& \frac{1}{\rho}\binom{c}{2} + O\left(\frac{1}{\rho^2}\right),
\end{align*}
by the inductive hypothesis for the first term on the right and using \eref{eq:T1U1} for the second term. By considering
\begin{align*}
U^{(k)}_{a,b,c} := {}& \inf\left\{t \geq 0:
{\mathcal D}^{(\infty)}_{a,b,c}(t) \in \{(a',b',0,0) : a',b'\in \mathbb{N}\}\right\}, \quad k=2,3,
\end{align*}
the cases $k=2,3$ are similar. $\mathbb{P}(I^{(k_1)} \cap I^{(k_2)}) = O(\rho^{-2})$ also follows from the fact that this event requires at least two transitions which are not recombinations during the time that $c > 0$.
\end{proof}
Should the coupling fail, we can say much about the sequence of events prior to $U^{(k)}_{a,b,c}$. Intuitively, the probability that \emph{more than} one transition other than recombinations occurs is $O(\rho^{-2})$. To make this precise we denote by ${\mathcal S}^{(k)}_{a,b,c}(t)$ the jump chain up to time $t$ of ${\mathcal C}^{(\rho)}$ if $k=1$ and of ${\mathcal D}^{(\infty)}$ if $k=2, 3$
\begin{lemma}
\label{lem:skeleton}
Let $\mathscr{S}_{a,b,c}$ denote the set of jump chains comprising sequences which start at $(a,b,c,c)$, end at the first entry of the form $(a',b',0,0)$, $a',b'\in\mathbb{N}$, and with all transitions corresponding to recombination events, except for possibly one transition. Then
\[
\mathbb{P}({\mathcal S}^{(k)}_{a,b,c}(U^{(k)}_{a,b,c}) \in \mathscr{S}_{a,b,c} \mid I^{(k)}) = 1 - O\left(\frac{1}{\rho}\right) \quad \text{as }\rho\to\infty, \quad k=1,2,3.
\]
\end{lemma}
\begin{proof}
The non-recombination event causing $I^{(k)}$ occurs at time $T^{(k)}_{a,b,c}$. Inspection of the generators \eref{eq:genrho} and \eref{eq:geninf2} shows that any further transition other than a recombination occurs with probability $O(\rho^{-1})$ during the time that $c > 0$.
\end{proof}
Recall that our purpose is to obtain the sampling distribution for ${\mathcal C}^{(\rho)}$. For successful couplings, this is easy to obtain since it is the same as that of ${\mathcal D}^{(\infty)}$ and hence ${\mathcal C}^{(\infty)}$; thus ${\mathcal C}^{(\rho)}\mid I^{(1)\complement}$ has the same sampling distribution as ${\mathcal D}^{(\infty)}\mid (I^{(2)} \cup I^{(3)})^\complement$. Even if the coupling fails, Lemmata \ref{lem:c=0} and \ref{lem:skeleton}, demonstrate that the behaviour of ${\mathcal C}^{(\rho)}$ is still predictable enough to recover its sampling distribution up to $O(\rho^{-2})$. Roughly [up to $O(\rho^{-2})$], \lemmaref{lem:skeleton} says: if there is an event that causes the coupling to fail then this is the \emph{only} non-recombination event in the failing process before $U^{(k)}_{a,b,c}$; by \lemmaref{lem:c=0}, if it has not failed by $U^{(k)}_{a,b,c}$ then the coupling will not fail after $U^{(k)}_{a,b,c}$.
The following theorem is proven in \citet{jen:son:2009:G}; however, the following proof gives a coherent, \emph{process-level} explanation for the result.
\begin{theorem}
\label{thm:coupling}
Expressing the sampling distribution for $({\mathcal C}^{(\rho)}_{a,b,c}(t) : t\geq 0)$ as in \eref{eq:main}, the first two terms are given by \eref{eq:zerothorder} and \eref{eq:firstorder}.
\end{theorem}
\begin{proof}
Denote by $q_{{\mathcal C}^{(\rho)}\mid I^{(1)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})$ the sampling distribution of the process ${\mathcal C}^{(\rho)}\mid I^{(1)}$. By Lemmata \ref{lem:c=0} and \ref{lem:skeleton}, this sampling distribution is obtained up to $O(\rho^{-1})$ by picking a pair of full fragments at random to coalesce, with the remaining $c-1$ fragments all undergoing recombination, and subsequently running the process as ${\mathcal D}^{(\infty)}_{a+c-1,b+c-1,0} (\stackrel{a.s.}{=} {\mathcal C}^{(\infty)}_{a+c-1,b+c-1,0})$. Hence,
\begin{align}
q_{{\mathcal C}^{(\rho)}\mid I^{(1)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) &= \sum_{i=1}^K\sum_{j=1}^L \frac{\binom{c_{ij}}{2}}{\binom{c}{2}} q_{{\mathcal C}^{(\infty)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}} - {\bfmath{e}}_{ij}) + O\left(\frac{1}{\rho}\right),\notag\\
&= \sum_{i=1}^K\sum_{j=1}^L \frac{\binom{c_{ij}}{2}}{\binom{c}{2}} q^A({\bfmath{a}}+{\bfmath{c}}_A-{\bfmath{e}}_i)q^B({\bfmath{b}}+{\bfmath{c}}_B-{\bfmath{e}}_j) + O\left(\frac{1}{\rho}\right). \label{eq:qI1}
\end{align}
(We can also ignore the possibility of mutation prior to $U^{(1)}_{a,b,c}$ since, by the same argument as in \lemmaref{lem:skeleton}, a mutation occurs during this phase with probability $O(\rho^{-1})$.) Similarly,
\allowdisplaybreaks
\begin{align}
q_{{\mathcal D}^{(\infty)}\mid I^{(2)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) &= \sum_{i=1}^K \frac{\binom{c_{i\cdot}}{2}}{\binom{c}{2}} q_{{\mathcal C}^{(\infty)}}({\bfmath{a}}+{\bfmath{c}}_A-{\bfmath{e}}_i,{\bfmath{b}}+{\bfmath{c}}_B,\bfmath{0}) + O\left(\frac{1}{\rho}\right),\notag\\
&= \sum_{i=1}^K \frac{\binom{c_{i\cdot}}{2}}{\binom{c}{2}} q^A({\bfmath{a}}+{\bfmath{c}}_A-{\bfmath{e}}_i)q^B({\bfmath{b}}+{\bfmath{c}}_B) + O\left(\frac{1}{\rho}\right),\label{eq:qI2}\\
q_{{\mathcal D}^{(\infty)}\mid I^{(3)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) &= \sum_{j=1}^L \frac{\binom{c_{\cdot j}}{2}}{\binom{c}{2}} q_{{\mathcal C}^{(\infty)}}({\bfmath{a}}+{\bfmath{c}}_A,{\bfmath{b}}+{\bfmath{c}}_B-{\bfmath{e}}_j,\bfmath{0}) + O\left(\frac{1}{\rho}\right),\notag\\
&= \sum_{j=1}^L \frac{\binom{c_{\cdot j}}{2}}{\binom{c}{2}} q^A({\bfmath{a}}+{\bfmath{c}}_A)q^B({\bfmath{b}}+{\bfmath{c}}_B-{\bfmath{e}}_j) + O\left(\frac{1}{\rho}\right), \label{eq:qI3}
\end{align}
and so, together with \lemmaref{lem:coupling}
and the observation that
\begin{multline*}
\mathbb{P}([I^{(2)} \cup I^{(3)}]^\complement)q_{{\mathcal D}^{(\infty)}\mid (I^{(2)} \cup I^{(3)})^\complement}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = q_{{\mathcal D}^{(\infty)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})\\ {}- \mathbb{P}(I^{(2)})q_{{\mathcal D}^{(\infty)}\mid I^{(2)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) - \mathbb{P}(I^{(3)})q_{{\mathcal D}^{(\infty)}\mid I^{(3)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) + O(\rho^{-2}),
\end{multline*}
we obtain
\begin{multline}
q_{{\mathcal D}^{(\infty)}\mid (I^{(2)} \cup I^{(3)})^\complement}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = \left[1 + \frac{2}{\rho}\binom{c}{2}\right]\bigg[q_{{\mathcal D}^{(\infty)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})\\
{}- \frac{1}{\rho}\binom{c}{2}q_{{\mathcal D}^{(\infty)}\mid I^{(2)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) -\frac{1}{\rho}\binom{c}{2}q_{{\mathcal D}^{(\infty)}\mid I^{(3)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})\bigg] + O\left(\frac{1}{\rho^{2}}\right). \label{eq:qnotI2notI3}
\end{multline}
The key decomposition is then
\begin{align}
q({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) = {}& \mathbb{P}(I^{(1)})q_{{\mathcal C}^{(\rho)}\mid I^{(1)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) + \mathbb{P}(I^{(1)\complement})q_{{\mathcal C}^{(\rho)}\mid I^{(1)\complement}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) \notag\\
= {}& \mathbb{P}(I^{(1)})q_{{\mathcal C}^{(\rho)}\mid I^{(1)}}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}})
{}+ \mathbb{P}(I^{(1)\complement})q_{{\mathcal D}^{(\infty)}\mid (I^{(2)}\cup I^{(3)})^\complement}({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) \label{eq:keydecomp}\\
={} & q_0({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) + \frac{1}{\rho}q_1({\bfmath{a}},{\bfmath{b}},{\bfmath{c}}) + O\left(\frac{1}{\rho^2}\right), \notag
\end{align}
using \eref{eq:failures}, \eref{eq:qI1}, \eref{eq:qI2}, \eref{eq:qI3}, and \eref{eq:qnotI2notI3}, with $q_0$, $q_1$ given by \eref{eq:zerothorder} and \eref{eq:firstorder}, respectively.
\end{proof}
\begin{remark}
It may be possible to use similar arguments to obtain a genealogical interpretation of the second-order term, $q_2$ in \eqref{eq:main}; for example, genealogies with \emph{two} events that cause the coupling to fail would surely contribute. However, as is clear from the expression for $q_2$ given in \citet{jen:son:2009:G, jen:son:2010:AAP}, this is not a simple endeavour and it is seems difficult to interpret some of the components of $q_2$.
\end{remark}
\subsection{A new ``loose-linkage'' coalescent process}
Equation \eref{eq:keydecomp} tells us that, up to $O(\rho^{-2})$, we can obtain the correct sampling distribution using the mixture
\[
\alpha[{\mathcal C}^{(\rho)}\mid I^{(1)}] + (1-\alpha)[{\mathcal D}^{(\infty)}\mid (I^{(2)}\cup I^{(3)})^\complement], \quad \alpha = \frac{1}{\rho}\binom{c}{2},
\]
provided $\alpha < 1$. The coupling used to prove \thmref{thm:coupling} demonstrates that we can \emph{define} a simple stochastic process for weakly correlated loci, ${\mathcal E}^{(\rho)}$, as follows, whose sampling distribution agrees with \eref{eq:zerothorder} and \eref{eq:firstorder} up to $O(\rho^{-2})$.\\
\fbox{
\begin{minipage}[c]{0.9\textwidth}
{\bf Algorithm to simulate ${\mathcal E}^{(\rho)}$, the \emph{loose-linkage coalescent}.}
\begin{enumerate}
\item With probability $\alpha$,
choose a pair uniformly at random from the $c$ full fragments to coalesce, and then choose uniformly from the chains in $\mathscr{S}_{a,b,c}$ compatible with $I^{(1)}$. Such chains are some permutation of a sequence corresponding to this sole coalescence and $c-1$ recombinations. Inter-event \emph{times} up to $U^{(1)}_{a,b,c}$ can be sampled according to the rates specified in \eref{eq:genrho}. Go to step 3.
\item Otherwise (w.p.\ $1-\alpha$), sample from ${\mathcal D}^{(\infty)}\mid (I^{(2)} \cup I^{(3)})^\complement$ up to time $U^{(2)}_{a,b,c}$ ($= U^{(3)}_{a,b,c}$), which can be achieved by running ${\mathcal D}^{(\infty)}$ as usual according to \eref{eq:geninf2} but banning transitions of the form $(a,b,c,d) \mapsto (a,b,c-1,d)$ and $(a,b,c,d) \mapsto (a,b,c,d-1)$. (The rates of these transitions still contribute to the overall rate governing inter-event times, however.) Go to step 3.
\item Beyond time $U^{(k)}_{a,b,c}$ ($k = 1$ in the first case above and $k=2$ in the second), construct the remainder of the process independently using $({\mathcal C}^{(\infty)}(t-U^{(k)}_{a,b,c}):t\geq U^{(k)}_{a,b,c})$ (with the appropriate starting configuration) back to the first time both loci have found a most recent common ancestor.
\end{enumerate}
\end{minipage}}
\\
\begin{figure}[t]
\centering
\setlength{\unitlength}{0.42cm}
\begin{picture}(12,9)(-2,0)
\thicklines
\put(0,0){\line(0,1){1}}
\put(2,0){\line(0,1){1}}
\put(4,0){\line(0,1){0.5}}
\put(7,0){\line(0,1){1.5}}
\put(0,1){\line(1,0){2}}
\put(1,1){\line(0,1){1}}
\put(3,0.5){\line(1,0){1}}
\put(3,0.5){\line(0,1){5.5}}
\put(6,1.5){\line(1,0){1}}
\put(6,1.5){\line(0,1){6.5}}
\put(0,2){\line(1,0){1}}
\put(0,2){\line(0,1){4}}
\put(0,6){\line(1,0){3}}
\put(1.5,6){\line(0,1){2}}
\put(1.5,8){\line(1,0){4.5}}
\multiput(0.1,0)(0,0.1){10}{\line(0,1){0.05}}
\multiput(2.1,0)(0,0.1){10}{\line(0,1){0.05}}
\multiput(4.1,0)(0,0.1){5}{\line(0,1){0.05}}
\multiput(7.1,0)(0,0.1){15}{\line(0,1){0.05}}
\multiput(0.1,0.9)(0.1,0){20}{\line(1,0){0.05}}
\multiput(1.1,1)(0,0.1){10}{\line(0,1){0.05}}
\multiput(4.1,0.5)(0.1,0){10}{\line(1,0){0.05}}
\multiput(5.1,0.5)(0,0.1){45}{\line(0,1){0.05}}
\multiput(7.1,1.5)(0.1,0){10}{\line(1,0){0.05}}
\multiput(8.1,1.5)(0,0.1){75}{\line(0,1){0.05}}
\multiput(1.1,2)(0.1,0){10}{\line(1,0){0.05}}
\multiput(2.1,2)(0,0.1){30}{\line(0,1){0.05}}
\multiput(2.1,5)(0.1,0){30}{\line(1,0){0.05}}
\multiput(3.6,5)(0,0.1){40}{\line(0,1){0.05}}
\multiput(3.6,9)(0.1,0){45}{\line(1,0){0.05}}
\thinlines
\multiput(-0.5,2.1)(0.5,0){19}{\line(1,0){0.25}}
\put(-2,2){\tiny $U^{(1)}_{0,0,4}$}
\put(-2,0.9){\tiny $I^{(1)}$}
\put(-1,1){\vector(1,0){0.9}}
\put(9.5,5.4){\raisebox{.5pt}{\textcircled{\raisebox{-.9pt} {3}}}}
\put(9,2.1){\line(1,0){0.25}}
\put(9,9){\line(1,0){0.25}}
\put(9.25,5.6){\line(1,0){0.125}}
\put(9.25,2.1){\line(0,1){6.9}}
\put(9.5,0.8){\raisebox{.5pt}{\textcircled{\raisebox{-.9pt} {1}}}}
\put(9,0){\line(1,0){0.25}}
\put(9,2){\line(1,0){0.25}}
\put(9.25,1){\line(1,0){0.125}}
\put(9.25,0){\line(0,1){2}}
\end{picture}
\hspace{35pt}
\begin{picture}(14.5,9)(-3,0)
\thicklines
\put(0,0){\line(0,1){2}}
\put(3,0){\line(0,1){3}}
\put(6,0){\line(0,1){0.5}}
\put(9,0){\line(0,1){1.5}}
\put(2,3){\line(1,0){1}}
\put(2,3){\line(0,1){3}}
\put(5,0.5){\line(1,0){1}}
\put(5,0.5){\line(0,1){6.5}}
\put(8,1.5){\line(1,0){1}}
\put(8,1.5){\line(0,1){5.5}}
\put(5,7){\line(1,0){3}}
\put(6.5,7){\line(0,1){1}}
\put(-1,2){\line(1,0){1}}
\put(-1,2){\line(0,1){4}}
\put(-1,6){\line(1,0){3}}
\put(0.5,6){\line(0,1){2}}
\put(0.5,8){\line(1,0){6}}
\multiput(0.1,0)(0,0.1){20}{\line(0,1){0.05}}
\multiput(3.1,0)(0,0.1){30}{\line(0,1){0.05}}
\multiput(6.1,0)(0,0.1){5}{\line(0,1){0.05}}
\multiput(9.1,0)(0,0.1){15}{\line(0,1){0.05}}
\multiput(0.1,2)(0.1,0){10}{\line(1,0){0.05}}
\multiput(1.1,2)(0,0.1){50}{\line(0,1){0.05}}
\multiput(2.6,7)(0,0.1){20}{\line(0,1){0.05}}
\multiput(2.6,9)(0.1,0){60}{\line(1,0){0.05}}
\multiput(6.1,0.5)(0.1,0){10}{\line(1,0){0.05}}
\multiput(7.1,0.5)(0,0.1){45}{\line(0,1){0.05}}
\multiput(9.1,1.5)(0.1,0){10}{\line(1,0){0.05}}
\multiput(10.1,1.5)(0,0.1){35}{\line(0,1){0.05}}
\multiput(3.1,3)(0.1,0){10}{\line(1,0){0.05}}
\multiput(4.1,3)(0,0.1){40}{\line(0,1){0.05}}
\multiput(1.1,7)(0.1,0){30}{\line(1,0){0.05}}
\multiput(7.1,5)(0.1,0){30}{\line(1,0){0.05}}
\multiput(8.6,5)(0,0.1){40}{\line(0,1){0.05}}
\put(0,1){\line(1,0){1.25}}
\put(3,1){\line(-1,0){1.25}}
\put(1.32,0.9){\tiny $\cancel{\phantom{\times}}$}
\thinlines
\multiput(-1.5,3.1)(0.5,0){24}{\line(1,0){0.25}}
\put(-3,3){\tiny $U^{(2)}_{0,0,4}$}
\put(-2,0.9){\tiny $I^{(2)}$}
\put(-1,1){\vector(1,0){0.9}}
\put(11,5.9){\raisebox{.5pt}{\textcircled{\raisebox{-.9pt} {3}}}}
\put(10.5,3.1){\line(1,0){0.25}}
\put(10.5,9){\line(1,0){0.25}}
\put(10.75,6.1){\line(1,0){0.125}}
\put(10.75,3.1){\line(0,1){5.9}}
\put(11,1.3){\raisebox{.5pt}{\textcircled{\raisebox{-.9pt} {2}}}}
\put(10.5,0){\line(1,0){0.25}}
\put(10.5,3){\line(1,0){0.25}}
\put(10.75,1.5){\line(1,0){0.125}}
\put(10.75,0){\line(0,1){3}}
\end{picture}
\caption{\label{fig:looselinkage}Sampling from the loose-linkage coalescent, ${\mathcal E}^{(\rho)}$, from an initial configuration $(0,0,4)$. Steps of the algorithm in the main text are denoted by circled numbers. \emph{Left}: Commence from step 1 (probability $\alpha$). Step 1 samples from an approximation to \mbox{${\mathcal C}^{(\rho)}\mid I^{(1)}$} which is correct to $O(\rho^{-2})$, back as far as time $U_{0,0,4}^{(1)}$. The jump chain sampled here is ${\mathcal S}^{(1)}_{0,0,4}(U^{(1)}_{0,0,4}) = ((0,0,4,4),(1,1,3,3),(1,1,2,2),(2,2,1,1),(3,3,0,0))$. Thereafter (step 3) the sample is constructed from ${\mathcal C}^{(\infty)}_{3,3,0}(t-U^{(1)}_{0,0,4})$. \emph{Right}: Commence from step 2 (probability $1-\alpha$). Step 2 samples from ${\mathcal D}^{(\infty)}_{0,0,4}(t)\mid (I^{(2)}\cup I^{(3)})^\complement$; a transition which would cause $I^{(2)}$ is banned. Thereafter (step 3) the sample is constructed from ${\mathcal C}^{(\infty)}_{4,4,0}(t-U^{(2)}_{0,0,4})$.}
\end{figure}
An example is shown in \fref{fig:looselinkage}. Simulation and inference under ${\mathcal E}^{(\rho)}$ should be straightforward, since its dynamics are little more complicated than those of a coalescent process with $\rho = \infty$. Unlike our diffusion process of \sref{sec:diffusion}, it does not seem easy to write down its sampling distribution to all orders in closed-form, since that of ${\mathcal D}^{(\infty)}\mid (I^{(2)}\cup I^{(3)})^\complement$ is not so obvious.
\section{Discussion}
\label{sec:discussion}
We have described two novel stochastic models of evolution for loosely linked, or weakly correlated, loci, using both diffusion- and coalescent-based arguments. As a consequence we have obtained deep insight into the simple form of the asymptotic sampling formula given by \eref{eq:zerothorder} and \eref{eq:firstorder}. Our diffusion model is based on a central limit theorem for density dependent population processes,
which may be viewed as a separation of the timescales $N^\beta$ and $N$ (in generations), for $0 < \beta < 1$, and pioneered in population genetics by \citet{nor:1975:SIAM}. This contrasts with most research in this area, which focuses on separating the timescales $N^0 = 1$ and $N$. Indeed, both diffusion \citep{eth:nag:1980, eth:nag:1988} and coalescent \citep{moe:1998:AAP30:493, wak:2008} limits of this latter regime have been studied in detail. It is also the setting of the ``loose linkage'' limit of \citet{eth:nag:1989}. Our usage of ``loose linkage'' therefore refers to a scaling intermediate between the usual Wright-Fisher diffusion and that of \citet{eth:nag:1989}. That the pioneering approach of \citet{nor:1975:SIAM} to investigate recombination does not seem to have been considered until now supports the observation that his work is ``somewhat neglected'' \citep{wak:2005}. It would also be of interest to find a coalescent-based analogue of these results
along the lines of \citet{moe:1998:AAP30:493}, or even a duality relationship in the manner of \citet{eth:gri:2009}.
For simplicity we have focused on a two-locus, finite-alleles, neutral model. Most of this article does not hinge heavily on these assumptions, and it should be relatively straightforward to extend our results to incorporate things like
natural selection and more sophisticated models of mutation.
\section*{Acknowledgments}
We gratefully acknowledge the support of the Isaac Newton Institute. Part of this work stemmed from discussions P.F.\ and Y.S.S.\ had during the 2010 program on ``Statistical Challenges Arising from Genome Resequencing.'' We also thank the generous support of the Simons Institute for the Theory of Computing. This work was completed while P.A.J.\ and Y.S.S.\ were participating in the 2014 program on ``Evolutionary Biology and the Theory of Computing.''
\bibliographystyle{imsart-nameyear}
|
\section{Introduction}
Central stars of planetary nebulae, or CSPNe, are evolved stages of low- and
intermediate-mass stars. They are very
hot, with surface temperatures between 20,000 and 200,000 K, and many of them
present signatures of strong stellar winds in their spectra. It is generally
believed that the mass loss observed in these stars is driven by radiation
pressure in spectral lines, the same mechanism thought to accelerate the winds
of hot massive stars and which is intrinsically unstable and leads to the
formation of density inhomogeneities usually called clumps \citep[see, for
example,][]{2008cihw.conf...17M,2002A&A...383.1113D,2003A&A...406L...1D,
2005A&A...437..657D,2007A&A...476.1331O,Sundqvist2013}.
CSPNe are the connection between the asymptotic giant branch (AGB) phase and the
white dwarfs (WDs) and are a unique opportunity to test stellar atmosphere,
structure and evolution models. Their winds have the potential to impact
subsequent stellar evolution and shape the planetary nebulae, which also owe
their ionization to the strong radiation fields of the central stars. Thus,
solid determinations of the photospheric
and wind parameters of CSPNe, as well as of their chemical abundances, can help
answering questions concerning stellar
evolution, in particular the connection among the different types of hydrogen
deficient central stars, the wind driving mechanism, and the surrounding
nebulae.
It is estimated that about 30 per cent of CSPNe are hydrogen deficient
\citep{Weidmann2011}. The main constituents of their atmospheres are, most
commonly, helium, carbon, oxygen, neon, and, often, also nitrogen. Among the
hydrogen deficient CSPNe, those evolving away from the AGB towards higher
temperatures (in the Hertzsprung-Russell diagram - HRD), showing strong carbon
and helium emission lines in their spectra, similar to massive Wolf-Rayet stars,
are named [WC] stars, which are further divided into early ([WCE]) and late
types ([WCL]), according to the ionization stages visible in their spectra. The
hydrogen deficient CSPNe positioned at the top of the white dwarf cooling track
in the HRD, exhibiting weak, or no wind features, in addition to absorption
lines of highly ionized helium, carbon, and oxygen, belong to the
PG1159 class. Due to the proximity of these types of
hydrogen deficient CSPNe in the $\log T_{\ast}-\log g$ diagram and also because
of similarities in their abundance patterns, they are thought to form an
evolutionary sequence, in which they would move away from the AGB and progress as [WC]
stars towards higher temperatures, evolving from [WCL] to [WCE] types as their
temperatures increase, until nuclear burning ceases and they enter the white
dwarf cooling track. Their luminosities and mass-loss rates decrease and their
winds reach the very high terminal velocities (up to 4000 km s$^{-1}$) typical
of [WC]-PG1159 and PG1159 stars
\citep{1991A&A...247..476W,2000A&A...362.1008G,2001A&A...367..983P}. Other
classes of H-deficient post-AGB objects exist, which will not be discussed here
\citep[see, for
example,][]{2007AJ....134.1380M,2010AIPC.1273..219T,2013MNRAS.430.2302T,
1998AeA...338..651R}.
He, C, O, and Ne abundances observed in [WC] and PG1159 stars - which, however,
vary greatly from one object to the other - resemble those of the intershell
region of AGB models. According to \citet{Werner2007a}, in the case of
PG1159 stars, the observed abundance intervals are: $X_{He}=0.30-0.85$,
$X_{C}=0.15-0.60$, and $X_{O}=0.02-0.20$. Abundance determinations for [WC] stars
are approximately in the intervals: $X_{He}=0.40-0.79$, $X_{C}=0.15-0.55$, and
$X_{O}=0.01-0.15$
\citep{1996A&A...312..167L,1997IAUS..180..114K,1997A&A...320...91K,
1998A&A...330..265L,1998A&A...330.1041K,2007ApJ...654.1068M}. Post-AGB
evolutionary models are able to predict the helium, carbon, and oxygen abundance
patterns
seen in PG1159 and [WC] stars, explaining the wide range of values seen by
differences in the
initial mass of the stars and on the number of thermal pulses suffered during
the AGB phase
\citep[see, for example,][and references therein]{Werner2007a}. Their hydrogen
deficiency is usually thought to be the result of late thermal pulses, which can
occur either at the tip of the AGB (in which case it is named an AGB final
thermal pulse or AFTP), or during the post-AGB evolutionary phase, as in the
very late thermal pulses (VLTP) of the born-again scenario of
\citet{1983ApJ...264..605I} (taking place after the star has entered the WD
cooling track in the HRD), or in the so called late thermal pulses (LTP)
occurring during the CSPN phase \citep[see][]{2001Ap&SS.275....1B,
2001Ap&SS.275...15H, 2005A&A...435..631A}. Each of these scenarios produces
H-deficient stars. However, the surface abundance patterns originated will
differ between the scenarios \citep[see, for example,][and references
therein]{2006PASP..118..183W}.
[WCE] CSPNe are thus the hottest among [WC]-type stars and are probably the
direct predecessors of [WC]-PG1159 and PG1159 stars, which are, in turn,
believed to evolve into DA and DO WDs \citep{Althaus2010}. Their
analysis can help establish constraints for the different post-AGB evolutionary
scenarios. The spectra shown by the hottest [WCE] stars are characterized by
lack of photospheric absorption lines, the presence of strong, broad emission
lines, UV P-Cygni profiles, and a paucity of lines in the optical region, in
which none of the few strong lines they present have P-Cygni profiles.
Essential diagnostic lines lay shortwards of Ly\,$\alpha$, in the far-UV range,
which was provided by FUSE (Far Ultraviolet Spectroscopic Explorer - 905$-$1187
$\mathrm{\AA}$ - \citet{2000ApJ...538L...1M}). In this region, these stars show
two conspicuous P-Cygni profiles: Ne\,{\sc vii} $\lambda$ $973.3$
$\mathrm{\AA}$ and O\,{\sc vi} $\lambda \lambda$ 1031.9, 1037.6 $\mathrm{\AA}$, as shown in Fig. \ref{FUSEspectra}. Complementary line diagnostics
are found in the spectral region between 1150 and 3200 $\mathrm{\AA}$, covered
by IUE (International Ultraviolet Explorer) and HST (Hubble Space Telescope)
spectrographs (Fig. \ref{UVspectra}). A detailed description of the UV spectra
of [WCE] CSPNe was provided by \citet{kelleretal2011PORT}.
\begin{figure}
\centering
\includegraphics[scale=0.48]{FUSE_wce_observedspectra.pdf}
\includegraphics[scale=0.48]{FUSE_wce_modelspectra.pdf}
\caption{Left panel: FUSE spectra of the [WCE] CSPNe NGC 6905, NGC 5189, Pb 6
and NGC 2867. The numerous narrow absorptions are due to interstellar H$_{2}$,
which affects the blue edge of the Ne\,{\sc vii} P-Cygni profile. Right panel: sample
synthetic spectra from our model grid, calculated for different values of
stellar temperature and two different mass-loss rates (in units of M$_{\odot}$
yr$^{-1}$). All models shown adopt $v_{\infty}=2500$ km s$^{-1}$ and models of
the same temperature adopt the same stellar radius. The observed and synthetic
spectra shown were convolved with a Gaussian of 0.2 $\mathrm{\AA}$ FWHM for
clarity.}\label{FUSEspectra}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.48\columnwidth]{UV_wce_observedspectra_nebids_c2.pdf}
\includegraphics[width=0.48\columnwidth]{plot_wce_models_uv_d2.pdf}
\caption{Left panel: HST/STIS G140L spectra of the [WCE] CSPNe NGC 6905, Pb 6
and Sand 3, and low resolution IUE spectra of NGC 5189 and NGC 2867. The dotted
vertical lines indicate the location of emission lines commonly seen in PNe.
\citet{Feibelman1996} identified the structure around 1320 $\mathrm{\AA}$ as
being a nebular [Mg V] line in the spectra of Sand 3. We, however, find evidence
of its stellar origin, as we discuss on Section \ref{sec:targets}. Right panel:
sample synthetic spectra from our model grid (same models as in Fig.
\ref{FUSEspectra}), between 1200 and 1750 $\mathrm{\AA}$, calculated for
different values of stellar temperature and two different mass-loss rates (in
units of M$_{\odot}$ yr$^{-1}$). The synthetic spectra shown were convolved with
a HST/STIS G140L instrumental LSF.}\label{UVspectra}
\end{figure}
In this paper, we present a comparative analysis of UV and far-UV spectra of the
hot [WCE]-type CSPNe NGC 2867, NGC 5189, NGC 6905, Pb 6, and Sand 3 (WO1 stars in the unified classification
scheme of \citet{1998MNRAS.296..367C}), prompted by the lack of a comprehensive, systematic analysis of [WCE]
CSPNe making use of the available high resolution spectra and of today's modern
stellar atmosphere codes. Our analysis includes many ionic species neglected in
previous works, thus allowing us to model, for the first time in [WCE] stars,
the Ne far-UV spectral lines, besides improving the overall match between model
and observations. The analysis was based on our extensive grid of synthetic
spectra \citep{kelleretal2011PORT} and numerous subsequent model
calculations for each individual object. A detailed analysis of
the central star of NGC 6905 was presented in Keller et al. (2011) and the results are now compared with those obtained for the other stars studied in this work,
placing the central star of NGC 6905 in the context of [WCE]
CSPNe. With the exception of Pb 6, all CSPNe studied here are classified as GW
Vir (=PG 1159$-$035) variables - a class of nonradial pulsating
variable H-deficient pre-White Dwarfs, showing variability $\la$0.3 mag \citep[see][and references therein]{Althaus2010}.
The paper is organized as follows: in Section \ref{sec:targets}, we describe the
sample objects and the data used in the analysis. Section \ref{sec:thegrid}
describes the synthetic spectra. In Section \ref{sec:SpecAnalyses}, we
describe the spectral analysis of the sample objects. Finally, we present our
conclusions in Section \ref{sec:conclusions}.
\section{Sample objects and data}
\label{sec:targets}
The objects studied here, the central stars of NGC 2867, NGC 5189, NGC 6905, Pb 6, and Sand 3, belong to the small group of [WCE] stars for which good
quality UV spectra exist: quality far-UV FUSE spectra exist for four of these
stars (NGC 2867, NGC 5189, NGC 6905, and Pb 6). However, only for NGC 6905 and Pb
6 there are complementary HST/STIS UV spectra. In the case of NGC 5189 and NGC
2867, we used complementary IUE low resolution spectra. Sand 3, on the other
hand, has excellent HST/STIS G140L and G230MB spectra, but no FUSE, which limits
the available line diagnostics, but its analysis is nonetheless interesting,
particularly for its spectral similarity to the other stars, especially to the
central star of NGC 6905. The available STIS spectra of Sand 3 were complemented
by low resolution IUE spectra between 1725 and 3300 $\mathrm{\AA}$.
Our sample objects are listed in Table \ref{objects}, along with radial
velocities ($v_{rad}$), distances, and nebular parameters (expansion velocity -
$v_{exp}$, angular diameter, and shape). Throughout this paper, the observed
spectra presented in the figures have been shifted to the rest frame of the
star, using the listed radial velocities. The utilized spectra are listed in
Table \ref{spectra}. IUE, HST/STIS, and FUSE spectra used in the analysis were retrieved from
MAST (Mikulski Archive for Space Telescopes - \url{http://archive.stsci.edu/}). NGC 2867, NGC 5189, and Pb 6 FUSE and Pb 6 and Sand 3 STIS spectra have not been
used in stellar atmosphere modelling before this work. NGC 6905's STIS spectra,
as well as the region of its FUSE spectra shortwards of 1000 $\mathrm{\AA}$, in which
the strong P-Cygni profile of the $\lambda$ $973.3$ $\mathrm{\AA}$ Ne\,{\sc vii} line
can be found, were first modelled by us and presented in
\citet{kelleretal2011PORT}. The objects were thus selected among the known
sample of hot [WCE] CSPNe for the availability of good UV spectra, most of them
not previously analysed.
\begin{table}
\caption{Sample objects.}
\label{objects}
\begin{tabular}{llllll}
\hline
Object & v$_{rad}$ [km s$^-1$] & Dist. [kpc] & Nebular Diam. [arcsec] &
v$_{exp}$ [km s$^-1$] & Shape \\
\hline
NGC 6905 & -8.4\tablenotemark{a} & 1.73\tablenotemark{e} ;
1.75\tablenotemark{f} ; 1.80\tablenotemark{g} & 43.3 $\times$
35.6\tablenotemark{d} & 43.5\tablenotemark{a} & Elliptical \\
NGC 5189 & -9.2\tablenotemark{a} & 0.54\tablenotemark{e} ;
0.55\tablenotemark{f} ; 0.70\tablenotemark{g} & 163.4 $\times$
108.2\tablenotemark{d} & 36.0\tablenotemark{a} & Quadrupolar \\
Pb 6 & 56.0\tablenotemark{b} & 4.38\tablenotemark{e} ;
4.42\tablenotemark{f} ; 4.00\tablenotemark{g} & 5.5\tablenotemark{e} & ? &
Unclassified \\
NGC 2867 & 14.4\tablenotemark{b} & 1.84\tablenotemark{e} ;
2.23\tablenotemark{f} ; 1.60\tablenotemark{g} & 14.4 $\times$
13.9\tablenotemark{d} & 18.9\tablenotemark{a} & Elliptical \\
Sand 3 & -92.5\tablenotemark{c} & ? & -- & -- & No detectable nebulosity
\\
\hline
\end{tabular}
$^{a}$\citet{Acker1992}; $^{b}$\citet{2013RMxAA..49...87P};
$^{c}$\citet{Feibelman1996}; $^{d}$\citet{Tylenda2003};
$^{e}$\citet{1992A&AS...94..399C}; $^{f}$\citet{Stanghellini2008};
$^{g}$\citet{Maciel1984}.
\end{table}
\begin{table}
\begin{center}
\caption{Spectral data sets analysed in this work.}
\label{spectra}
\begin{tabular}{llllllll}
\hline
Instrument & Data ID & Date & Resol. [$\mathrm{\AA}$] & Aperture [''] & Range
[$\mathrm{\AA}$] & comments \\
\hline
\textbf{NGC 6905} &&&&&&\\
FUSE & A1490202000 & 2000 Aug 11 & $\sim$0.06 & 30$\times$30 &
905$-$1187 & \\
STIS + G140L & O52R01020 & 1999 Jun 29 & $\sim$1.20 & 52$\times$0.5 &
1150$-$1730 & \\
STIS + G230L & O52R01010 & 1999 Jun 29 & $\sim$3.15 & 52$\times$0.5 &
1570$-$3128 & \\
\textbf{NGC 5189} &&&&&&\\
FUSE & S6013001000 & 2002 Feb 28 & $\sim$0.06 & 30$\times$30 &
905$-$1187 & \\
IUE & LWR07171 & 1980 Mar 12 & $\sim$7.0 & 10$\times$20 &
1850$-$3300 & \\
IUE & SWP08219 & 1980 Mar 12 & $\sim$6.0 & 10$\times$20 &
1150$-$2000 & \\
IUE & SWP06327 & 1979 Aug 30 & $\sim$6.0 & 10$\times$20 &
1150$-$2000 & off star\\
IUE & LWR01985 & 1978 Aug 05 & $\sim$7.0 & 10$\times$20 &
1850$-$3300 & off star\\
IUE & SWP02206 & 1978 Aug 05 & $\sim$6.0 & 10$\times$20 &
1150$-$2000 & off star\\
IUE & SWP25365 & 1985 Mar 05 & $\sim$6.0 & 10$\times$20 &
1150$-$2000 & off star\\
IUE & LWP05456 & 1985 Mar 05 & $\sim$7.0 & 10$\times$20 &
1850$-$3300 & \\
IUE & LWP05457 & 1985 Mar 05 & $\sim$7.0 & 10$\times$20 &
1850$-$3300 & off star\\
IUE & SWP25363 & 1985 Mar 05 & $\sim$0.2 & 10$\times$20 &
1150$-$2000 & bad extraction\\
IUE & SWP25364 & 1985 Mar 05 & $\sim$6.0 & 10$\times$20 &
1150$-$2000 & \\
IUE & SWP41913 & 1991 Jun 24 & $\sim$6.0 & 10$\times$20 &
1150$-$2000 & off star\\
\textbf{Pb 6} &&&&&&\\
FUSE & Z9111101000 & 2003 Feb 20 & $\sim$0.06 & 30$\times$30
& 905$-$1187 & \\
STIS + G140L & OBRZ23010 & 2012 Apr 25 & $\sim$1.20 & 52$\times$0.2 &
1150$-$1730 & \\
STIS + G230L & OBRZ23020 & 2012 Apr 25 & $\sim$3.15 & 52$\times$0.2 &
1570$-$3180 & \\
STIS + G230L & OBRZ23030 & 2012 Apr 25 & $\sim$3.15 & 52$\times$0.2 &
1570$-$3180 & \\
\textbf{NGC 2867} &&&&&&\\
FUSE & Z9110901000 & 2003 Feb 22 & $\sim$0.06 & 30$\times$30
& 905$-$1187 & \\
IUE & SWP05234 & 1979 May 14 & $\sim$6.0 & 10$\times$20 & 1150$-$2000 & noisy
\\
IUE & SWP05215 & 1979 May 12 & $\sim$6.0 & 10$\times$20 & 1150$-$2000 & \\
IUE & LWR04510 & 1979 May 12 & $\sim$7.0 & 10$\times$20 & 1850$-$3300 &\\
IUE & LWR04518 & 1979 May 14 &$\sim$7.0 & 10$\times$20 & 1850$-$3300 &\\
IUE & SWP08984 & 1980 May 12 & $\sim$0.2 & 10$\times$20 & 1150$-$2000 & bad
extraction\\
IUE & SWP52948 & 1994 Nov 30 & $\sim$0.2 & 10$\times$20 & 1150$-$2000 & bad
extraction\\
IUE & LWR07736 & 1980 May 12 & $\sim$0.2 & 10$\times$20 & 1850$-$3300 & bad
extraction\\
\textbf{Sand 3} &&&&&&\\
STIS + G140L & O4XH01030 & 1998 Sep 03 & $\sim$1.20 & 52$\times$2 &
1138$-$1725 & \\
STIS + G230MB & O4XH01020 & 1998 Sep 03 & $\sim$0.33 & 52$\times$2 &
2716$-$2872 & \\
IUE & SWP53882 & 1995 Feb 10 & $\sim$6.0 & 10$\times$20 &
1150$-$1975 & \\
IUE & LWR05765 & 1979 Oct 06 & $\sim$7.0 & 10$\times$20 &
1860$-$3300 & \\
\hline
\end{tabular}
\end{center}
\footnotesize{SWP53882 is only used in the interval 1725$-$1851 $\mathrm{\AA}$.}
\normalsize
\end{table}
\subsection{Description of the utilized data}
\subsubsection{FUSE data}
The FUSE instrument included four channels (LiF1, LiF2, SiC1, SiC2), each
divided into two segments (A and B), collecting light simultaneously. Each
channel-segment combination spans a fraction of FUSE's nominal wavelength
coverage, with some overlap among them. All FUSE spectra used in this work were
obtained through the $30''\times30''$ aperture, in time-tag mode.
We used the CalFUSE v3.2.2 pipeline for data reduction, along with the FUSE IDL
tools `CF\_edit' and `FUSE\_register'. We screened the observations for bad time
intervals, and checked for possible drifting of the target out of the aperture.
The spectra of the different segments were then extracted with the
`cf\_extract\_spectra' routine. Faulty regions of the spectra, such as those
close to the detectors' edges and the region of the LiF 1B detector affected by
the artefact known as `worm' (due to a shadow thrown on the detector by the
instrument itself) were removed and the spectra of the different segments,
co-added weighted by their errors.
In the case of CSPN Pb 6, continuum flux levels differ between FUSE and HST/STIS
spectra, with the former being 40 per cent higher. Such divergences in flux
levels are often related to the positioning of the target into the instrument's
aperture or difficulties in extracting the background in low signal-to-noise
data. A comparison between HST G140L and IUE low resolution (SWP29857) spectra
of Pb 6 shows agreement in continuum levels within 10 per cent. Flux levels also agree between
G140L and G230L spectra of Pb 6. According to \citet{1998stis.rept...20B}, the STIS
$52''\times0.2''$ aperture is calibrated to a few percent accuracy, with the repeatability for point
sources being at worse 10 per cent, with a RMS of 4.5 per cent. We thus have scaled the flux level
of the FUSE spectrum of Pb 6 down by dividing it by 1.4 to match that of the
STIS spectrum.
The FUSE spectra of the central stars studied here present some strong and
narrow emissions that do not seem to originate in the atmospheres of these
stars, for they are much narrower than their typically broad wind lines. Some of
these emissions are due to the planetary nebulae and some are due to airglow or
scattered sun light, as verified by comparing the spectra obtained during the
orbital night with the composite of all data. Except for NGC 2867, for which the
signal to noise ratio of the available FUSE spectra allowed us to use night only
events, thus minimizing airglow contamination, both orbital night and day
time-tag events were used in the stellar spectra of the sample objects.
The central stars of NGC 2867 and NGC 5189 have 3 FUSE observations each. We have used those with the
highest signal-to-noise ratio. While the plot previews available
at the MAST website show some variation of flux levels and line profiles between different data sets, the careful treatment of the data, as
described above, aiming at mitigating the effects of channel drift, airglow, faulty centroid positioning, worms
and dead zones, strongly diminishes the disagreements between the spectra obtained in different observations,
which are now on the same level as those seen comparing spectra from the same observation, but from different channels, suggesting that instrumental effects plus others such as airglow (which can be minimized by the use of night only time-tag events, but is never completely absent) may be responsible
for the remaining small variations in line profiles and flux levels.
\subsubsection{IUE data}
Despite the fact that MAST lists all IUE spectra for NGC 5189 under the category
70, which is described there as observation of nebula plus central star, our
analysis of the spectra indicate that many of the exposures actually missed the
central star, as in the case of SWP25363 - the only high resolution IUE spectrum
available for NGC 5189. In the 2D high resolution SWP25363 spectral image,
spatially extended nebular emission lines and only a very faint continuum are
visible. The continuum is not centred, but at the edge of the illuminated
region, which is evidence that flux was partially missed. The pipeline seems to
have misplaced the extraction slit and part of the resulting net flux is below
zero. Nonetheless, the spectral resolution of 0.2 $\mathrm{\AA}$ allows us to
estimate the width of the He\,{\sc ii} $\lambda$ 1640.4 $\mathrm{\AA}$ line as $\approx$0.6
$\mathrm{\AA}$, consistent with a nebular origin. We
thus cannot establish from the echelle spectrum whether there is also stellar He\,{\sc ii} $\lambda$ 1640.4 $\mathrm{\AA}$ emission. Fig. \ref{subtraction} shows IUE
SWP25364 and SWP25365 low-resolution spectra, which were taken at the same
pointing coordinates and with the large aperture (10'' $\times$ 20''),
but at different position angles. The latter seems to have at least partially
missed the central star. The difference between the two spectra exposes the
nebular origin of the C\,{\sc iii]} $\lambda$ 1909 $\mathrm{\AA}$ line and indicates
that at least part of the He\,{\sc ii} $\lambda$ 1640.4 $\mathrm{\AA}$ emission seen in
SWP25364 is of nebular origin. Therefore, we will regard NGC 5189's He\,{\sc ii}
$\lambda$ 1640.4 $\mathrm{\AA}$ emission as an upper limit to the stellar line
and leverage on other spectral lines to constrain the stellar parameters.
Since there are no high-resolution spectra of NGC 5189's central star in the
region between 1200 and 3200 $\mathrm{\AA}$, we rely on the low resolution
data. For the present analysis of NGC 5189, we use SWP08219 and
LWR07171, for being better centred and less noisy than the SWP25364 and
LWP05456 spectra, respectively.
The available IUE low resolution spectra of NGC 2867 seem to be dominated by
nebular emissions, such that P-Cygni profiles cannot be distinguished. The
inspection of the spectral images of SWP05215 and LWR04510 shows well defined,
well centred and narrow continua. Since only the strongest emission lines in the
SWP05215 spectrum appear saturated and because its continuum has a higher signal than that of the noisy SWP05234 spectrum, we chose to show the former instead of
the latter in this work, for it allows a more precise determination of the slope
of the continuum, which will be used in the normalization of the observed
spectra and to ascertain the colour excess. We will not use it for line strength
analysis, except as an upper limit for stellar emission in some weaker lines.
The existing high resolution IUE spectra of this star exhibit again very faint
continua (in the two-dimensional spectra) and negative flux values (in the extracted spectra), indicating
again extraction problems. No obvious stellar lines are apparent. Narrow and
strong C\,{\sc iv} $\lambda \lambda$ $1548.2$, $1550.8$ $\mathrm{\AA}$, He\,{\sc ii} $\lambda$
1640.4 $\mathrm{\AA}$ and C\,{\sc iii]} $\lambda$ 1909 $\mathrm{\AA}$ nebular emission
lines are, however, readily discernible in the high resolution spectra.
\begin{figure}
\centering
\includegraphics[scale=0.435]{subtracted_spectra_fixedforpaperII_final.pdf}
\caption{SWP25364 (red) and SWP25365 (black) low-resolution IUE spectra. The
result of subtracting SWP25365 from
SWP25364 is shown in green.}\label{subtraction}
\end{figure}
\subsubsection{HST data}
HST/STIS gratings are designed for spatially resolved spectroscopy using a long
slit, exploiting HST's high spatial resolution capabilities. The HST/STIS
spectra of CSPNe NGC 6905, Pb 6, and Sand 3 thus minimize nebular contamination
of the stellar extracted spectra - in comparison with IUE spectra, for example.
The inspection of the 2D G140L and G230L spectra of NGC 6905 and Pb 6 reveals
extended nebular emission around $\approx$1909 (very faint in NGC 6905),
1550 and 1640 $\mathrm{\AA}$, as seen on the logarithmic scale plots in the
supplementary figures, online. Other faint extended emissions are seen in
the 2D G140L spectrum of Pb 6 around $\approx$1240, 1400 and 1485 $\mathrm{\AA}$.
The 2D Sand 3 G140L spectrum does not exhibit evidence of nebular emissions, in
line with the idea of its nebula having already dissipated \citep{Barlow1982}.
In neither 2D STIS spectra there is evidence of extended nebular emission close to the structure
associated by \citet{Feibelman1996} with nebular [Mg\,{\sc v}] $\lambda \lambda$
$1324.5$ $\mathrm{\AA}$ in
the low resolution IUE spectra of Sand 3 and marked in Fig. \ref{UVspectra}, even though the extracted stellar spectra
of all three objects observed with STIS show the feature,
which, together with the evidence of Sand 3's nebula having dispersed, points towards a stellar origin for this spectral feature.
Similarly, \citet{Feibelman1996a} suggested the nebular He\,{\sc ii} $\lambda$ 2253 $\mathrm{\AA}$ and numerous [Ne\,{\sc v}] lines between
2235 and 2265 $\mathrm{\AA}$ seen in the high resolution IUE spectra of NGC 6905 as
the sources of the wide emission feature seen in lower resolution spectra of
this region. The origin of this spectral feature, however, seems to be at least partially stellar, since
the strong feature appears not only in the low resolution IUE spectra of Sand 3
- despite its dispersed nebula not producing a detectable C\,{\sc iii]} $\lambda$ 1909
$\mathrm{\AA}$ emission - but also in the 1D STIS spectra of NGC 6905 and Pb 6,
whose 2D G230L spectra do not show evidence of extended nebular emission in this region.
The agreement between the medium resolution
($\approx$0.33 $\mathrm{\AA}$) HST/STIS G230MB observed spectrum of Sand 3's O\,{\sc v}
$\lambda \lambda$ $2781.0$, $2787.0$ $\mathrm{\AA}$ line and the low resolution IUE
spectrum of the same star indicates that this strong and wide line is from stellar origin. This was also verified from the inspection of the 2D G230MB spectrum
of Sand 3 and of the 2D G230L spectra of NGC 6905 and Pb 6, which do not show
extended emissions in this region.
\section{Modelling}
\label{sec:thegrid}
\subsection{Stellar models}
The spectra of these very hot, H-poor CSPNe are dominated by broad, strong
emission lines, which are signatures of their dense, high velocity stellar
winds. Spectra of hot objects with detectable winds require modelling with
non-LTE stellar atmosphere codes that account for expanding atmospheres. In this
work, we used models computed with CMFGEN
\citep{1998ApJ...496..407H,2005ApJ...627..424H}, which is a non-LTE,
non-hydrodynamic code that considers a stationary mass loss in a spherically
symmetric geometry, line blanketing (through the super-level approximation) and
wind clumping (through the filling factor approach). The detailed workings of
the code are described in the references above. The fundamental stellar
parameters include stellar temperature ($T_{\ast}$) and radius ($R_{\ast}$),
defined at an optical depth of $\tau=$20, mass-loss rate ($\dot{M}$), terminal
velocity of the wind ($v_{\infty}$), velocity law, and elemental abundances. We
assume a standard velocity law, $v(r)=v_{\infty}(1-r_{0}/r)^{\beta}$, where
$r_{0} \approx R_{\ast}$ and $\beta = 1$, and an exponential law for the clumping filling factor
(which is the fraction of the wind volume occupied by clumps), given by
$\textit{f}(v) = \textit{f}_{\infty} + (1 -
\textit{f}_{\infty})\exp{(-v(r)/v_{clump})}$, where $v_{clump}=200$ km s$^{-1}$
is the velocity beyond which the wind is no longer smooth and the clumping increases until the terminal value of $\textit{f}_{\infty}=0.1$ is
reached.
The analysis leverages on the extensive grid of synthetic spectra computed by
\citet{kelleretal2011PORT}, covering parameter values typical of H-poor CSPNe
and approximately following the evolutionary tracks from
\citet{2006A&A...454..845M}.
The grid allows for a uniform and systematic comparison of the spectral
features and facilitates line identification and the determination of the main
stellar parameters. The grid models adopt a constant abundance pattern:
X$_{\mathrm{C}}=0.45$, X$_{\mathrm{He}}=0.43$, X$_{\mathrm{O}}=0.08$,
X$_{\mathrm{N}}=0.01$, X$_{\mathrm{Ne}}=0.02$ in mass fractions, an iron
abundance a factor of $100$ lower than the solar value, and solar abundances for
all other elements included into the calculations. Within the parameter space of
[WCE] CSPNe, the grid models have temperatures of $\approx$ 100,000, 125,000,
150,000, 165,000, and 200,000 K. The detailed emerging flux was calculated
adopting a varying microturbulence velocity described by
$\xi(r)=\xi_{min}+(\xi_{max}-\xi_{min})v(r)/v_{\infty}$, with $\xi_{min}=10$ km
s$^{-1}$ near the photosphere and $\xi_{max}=50$ km s$^{-1}$. The grid models
were computed including many ionic species not considered in previous analyses
and are described in \citet{kelleretal2011PORT},
\citet{Keller2012a,KellerESOinpress}.\footnote{The grid covering the parameter
regime of the [WC] CSPNe is available on-line at
\url{http://dolomiti.pha.jhu.edu/planetarynebulae.html}, where the user will
find tables listing the available models and individual pages for every model,
with information about the stellar parameters adopted, the ions considered and
the synthetic spectra, as well as figures comparing the different models and
identifying the spectral lines predicted.} Additional ad-hoc models were
computed for each object with varying elemental abundances and microturbulence
velocities, and including additional less abundant ions. The ions included
in the final best-fitting models, together with the number of levels and
superlevels adopted in the calculations, are given in Table \ref{tab:ions}.
\begin{table}
\begin{center}
\scriptsize
\caption{Ion superlevels and levels of the best-fitting final models for the
central stars of NGC 6905, Sand 3, Pb 6, NGC 2867, and NGC 5189.}
\begin{tabular}{cccccccccccc}
\hline
Element & I & II & III & IV & V &
VI & VII & VIII & IX & X
& XI \\
\hline
He & 40,45 & 22,30 & 1 & & &
& & & &
& \\
C & & & & 49,64 & 1 &
& & & &
& \\
N & & & & & 13,21 &
1 & & & &
& \\
O & & & &29,48\tablenotemark{a}& 58,163
& 41,47 & 1 & & &
& \\
Ne & & & &45,355\tablenotemark{a}& 37,166
& 36,202 & 38,182 & 24,47 & 1 &
& \\
Na & & & & & &
52,452 & 37,251 & 72,214 & 27,71 & 1
& \\
Mg & & & & & 43,311 &
46,444 & 54,476 & 1 & &
& \\
Si & & & &22,33 & 33,71\tablenotemark{a}&
33,98\tablenotemark{a}& 1\tablenotemark{a}& &
& & \\
P & & & & & 16,62 &
1 & & & &
& \\
S & & & & & &
28,58 & 1 & & &
& \\
Ar & & & & & &
30,205 & 33,174 & 57,72 & 1 &
& \\
Ca & & & & & &
47,108\tablenotemark{a}& 55,514\tablenotemark{a}& 54,445\tablenotemark{a}&
35,367\tablenotemark{a}& 31,79\tablenotemark{a}& 1\tablenotemark{a} \\
Fe & & & & & &
44,433\tablenotemark{a}& 41,252 & 53,324 & 52,490
& 43,210 & 1 \\
Co & & & & & &
& 45,1000 & 50,1217 & 24,355 & 1
& \\
Ni & & & & & &
& 37,308 & 113,1000 & 75,1217 & 1
& \\
\hline
\end{tabular}
\label{tab:ions}
\tablenotetext{a}{These ionic species are not present in the best-fitting final
models for the central stars of NGC 5189, Pb 6, and NGC 2867.}
\normalsize
\end{center}
\end{table}
\subsection{Effects of the interstellar medium}
In order to separate stellar features from interstellar absorptions due to
neutral and molecular hydrogen along the line of sight, which greatly affect the
far-UV spectra of these objects, we modelled them as described in
\citet{Herald2002,2004ApJ...609..378H,Herald2004a}, assuming a temperature of the interstellar gas of 100 K, which is a typical value for the ISM, and determined the column densities. The derived interstellar parameters
are given in Table \ref{tab:is}, where $N(H$\,{\sc i}$)$ and $N(H_{2})$ are the neutral
and molecular hydrogen column densities, respectively. Fig. \ref{sand3Lyalpha}
shows the interstellar Ly\,$\alpha$ absorption seen in the observed spectra of
Sand 3 and illustrates the fact that accounting for atomic hydrogen ISM
absorption can affect the modelling of the N\,{\sc v} $\lambda \lambda$ 1238.8, 1242.8
$\mathrm{\AA}$ line. We did not attempt to model the ISM Ly\,$\alpha$
absorption in low resolution IUE spectra of CSPNe NGC 2867 and NGC 5189.
\begin{figure}
\centering
\includegraphics[scale=0.5]{sand3_Habs_0b_newnorm.pdf}
\caption{The interstellar Ly\,$\alpha$ absorption seen in the observed spectra
of Sand 3
(continuous black line) is compared to our model, to which we applied the effect
of different values of neutral hydrogen column density.
The model shown has $T_{\ast}=150$ kK, $R_{t}=$9.4 $R_{\odot}$ and
$X_{N}=0.07$}\label{sand3Lyalpha}
\end{figure}
\begin{table}
\begin{center}
\scriptsize
\caption{Interstellar parameters adopted in the modelling.}
\begin{tabular}{lccccc}
\hline
Star & $\log N(H$\,{\sc i}$)$\tablenotemark{b} & $\log N(H_{2})$\tablenotemark{c} &
$T_{gas}$ [K]\tablenotemark{d} & $E_{B-V}$ [mag] & R$_{v}$\\
\hline
Pb 6 & 21.50 & 20.20 & 100 & 0.25$\pm^{0.05}_{0.05}$ & 3.1\\
NGC 5189 & $-$ & 20.30 & 100 & 0.23$\pm^{0.07}_{0.03}$ & 3.1\\
NGC 2867 & $-$ & 19.90 & 100 & 0.18\tablenotemark{e} & 3.1\\
NGC 6905\tablenotemark{a} & 21.00 & 19.50 & 100 & 0.17$\pm^{0.13}_{0.07}$ & 3.1
\\
Sand 3 & 21.55 & $-$ & 100 & 0.40$\pm^{0.05}_{0.05}$ & 3.1\\
\hline
\end{tabular}
\label{tab:is}
\tablenotetext{a}{From \citet{kelleretal2011PORT}.}
\tablenotetext{b}{From Ly\,$\alpha$. $N(H$\,{\sc i}$)$ is given in units of
cm$^{-2}$.}
\tablenotetext{c}{From the absorption lines in the FUSE range. $N(H_{2})$ is
given in units of cm$^{-2}$.}
\tablenotetext{d}{Assumed.}
\tablenotetext{e}{Derived from the extinction quantity taken from
\citet{2002ApJS..138..279M}.}
\normalsize
\end{center}
\end{table}
\section{Spectral analyses}
\label{sec:SpecAnalyses}
As the first step in the spectral analysis of the stellar lines, we estimated
the terminal velocity of the wind from the short-wavelength edge of the strong
P-Cygni profile of C\,{\sc iv} $\lambda$ 1548.2 $\mathrm{\AA}$ and also from the long
wavelength component of the O\,{\sc vi} $\lambda \lambda$ $1031.9$, $1037.6$
$\mathrm{\AA}$ doublet whenever both or any of these lines were available,
assuming a turbulence of $10$ per cent the terminal velocity of the wind.
The spectra of [WCE] CSPNe are particularly poor in spectral lines of different
ionization stages of the same element, and for this reason, the absence or
presence of structures due to ions of different ionization potentials and the
overall appearance of the spectra are also used to constrain temperature. As was
shown by \citet{kelleretal2011PORT}, oxygen is one of the few elements producing
strong lines of different ionization stages in the spectra of [WCE] CSPNe, with
several O\,{\sc vi} and two O\,{\sc v} lines visible in the synthetic spectra throughout the
temperature range of [WCE] CSPNe. As found by \citet{kelleretal2011PORT}, one
important aspect about these O\,{\sc v} lines, which can influence the determination of
the stellar temperature, is their considerable increase in strength, in the
synthetic spectra, when less abundant ions are included into the calculation,
especially Mg, Na, Co, and Ni. These ions are not included in the grid models,
because of computational limitations, but we include them in our final
best-fitting models, as we show below.
The most important line diagnostics we rely upon when constraining the surface
temperatures of these stars are the Ne\,{\sc vii} $\lambda$ 973.3 $\mathrm{\AA}$ and O\,{\sc vi} $\lambda \lambda$ 1031.9, 1037.6 $\mathrm{\AA}$ P-Cygni profiles and the O\,{\sc v}
$\lambda$ $1371.3$ $\mathrm{\AA}$ and $\lambda \lambda$ $2781.0$, $2787.0$
$\mathrm{\AA}$ lines. The Ne\,{\sc vii} $\lambda$ 973.3 $\mathrm{\AA}$ line is a useful
temperature diagnostic, because its synthetic profile is substantially weakened
by the decrease in temperature in the [WCE] parameter regime and because it
shows little sensitivity to mass loss at model temperatures of 100,000 and
125,000 K, in the parameter regime covered by our model grid (Fig.
\ref{FUSEspectra}). At these temperatures, it also shows little sensitivity to
neon abundance, as shown by \citet{kelleretal2011PORT}. A similar behaviour with
temperature is observed in the synthetic O\,{\sc vi} $\lambda \lambda$ 1031.9, 1037.6
$\mathrm{\AA}$ doublet, whose strength also increases towards models with
T$_{\ast}=$165,000 K and which is not particularly sensitive to variations in
mass loss at model temperatures of 100,000 and 125,000 K. Thus, the strong Ne\,{\sc vii} $\lambda$ 973.3 $\mathrm{\AA}$ and O\,{\sc vi} $\lambda \lambda$ 1031.9, 1037.6
$\mathrm{\AA}$ P-Cygni profiles observed in all the objects with FUSE spectra
shown here, immediately point towards models of T$_{\ast} = $150,000 K or
165,000 K, with the latter allowing for even stronger line profiles,
concomitantly with good fits to the remaining lines.
One of the most striking differences among the observed spectra of the studied
stars, which are otherwise very similar, is the strength of the O\,{\sc v} lines. The
central star of NGC 6905 and Sand 3 show much stronger lines than is seen in the
spectra of NGC 5189 and of Pb 6, or even than is present on the nebula-dominated
IUE spectra of NGC 2867. This characteristic, along with the line
appearance, can further constrain temperature: the appearance of the O\,{\sc v}
$\lambda$ $1371.3$ $\mathrm{\AA}$ line profile also excludes models with
T$_{\ast} =$100,000 and 125,000 K, since none of the observed line profiles show
the deep P-Cygni absorption components that are seen in these models. Also,
although the grid models with T$_{\ast} = $150,000 K and 165,000 K show weak O\,{\sc v}
lines for mass-loss rates that allow reasonable fits of the overall observed
spectra, their synthetic profiles do become much stronger through the addition
of further ions into the calculations, with the T$_{\ast} = $150,000 K models
producing O\,{\sc v} profiles considerably stronger than models with T$_{\ast} =
$165,000 K. We thus concluded that the central star of NGC 6905 and Sand 3 are
cooler than the other three objects and best fit by models of T$_{\ast} =
$150,000 K, while the central stars of NGC 2867, NGC 5189, and Pb 6 best-fitting
models have T$_{\ast} = $165,000 K.
Hotter models, with T$_{\ast} = $200,000 K, were also computed by us. However,
they do not fit well the observed spectra: at the same time, the models show too
weak O\,{\sc vi} $\lambda \lambda$ 1031.9, 1037.6 $\mathrm{\AA}$ lines and too strong
He\,{\sc ii} $\lambda$ 1640.4 $\mathrm{\AA}$ and C\,{\sc iv} $\lambda\lambda$ 1107.6, 1107.9
$\mathrm{\AA}$, $\lambda\lambda$ 1168.8, 1169.0 $\mathrm{\AA}$ and
$\lambda \lambda$ $1548.2$, $1550.8$ $\mathrm{\AA}$ lines, such that the
discrepancy is too big to be solved by different abundance patterns. The 200,000
K models also predict Ne\,{\sc vii} $\lambda$ 973.3 $\mathrm{\AA}$ lines with shallower
and narrower absorption components than what is seen in the observations and no
O\,{\sc v} emission lines.
Another factor that influences the appearance of the spectral lines in these
stars is how dense their wind is. \citet{1989A&A...210..236S} have introduced
the so called transformed radius, defined as
$R_{\mathrm{t}}=R_{\ast}[(v_{\infty}/2500\ \mathrm{km}\ \mathrm{s}^{-1}) /
(\dot{M}/10^{-4}\ \mathrm{M}_{\odot}\mathrm{yr}^{-1})]^{2/3}$. Models of the
same transformed radius that also have identical temperature, composition, and
terminal velocity - that is, models where $R_{\ast} / \dot{M}^{2/3}$ is the same
- are known to have similar ionization structure and temperature stratification
and to lead to very similar wind line spectra \citep{1993A&A...274..397H}. In
this work, we will constrain the transformed radius rather than the mass-loss
rate, because deriving mass-loss from $R_{t}$ depends on a measurement of the stellar radius, which depends on the distance and distances to Galactic CSPNe
are commonly not well known. On top of this, the absence of strong absorption
lines not contaminated by wind emissions in the spectra of [WCE] stars prevents
us from measuring $\log g$.
The overall strength of all available wind lines was used by us to constrain the
transformed radius. Among them, the He\,{\sc ii} $\lambda$ 1640.4 $\mathrm{\AA}$ -
whenever not strongly affected by nebular emission - is especially useful due to
its little sensitivity to temperature in the parameter regime covered by the
model grid for [WCE] stars. The behaviour of the synthetic spectral lines
apparent in the spectra of [WCE] stars with stellar parameters was extensively
described in \citet{kelleretal2011PORT}.
After having used the grid of synthetic spectra to constrain stellar temperature
and transformed radius, we then extended the analysis by exploring various
abundances for the main elements and different microturbulence velocities, as
well as by including additional elements into the calculations. The temperature steps of our model grid, in the [WCE] regime,
are small enough that subsequent models do not show
strong variations of line diagnostics. That is the reason for the non-homogeneous temperature
intervals in our model grid. In our ad-hoc analysis, temperature was therefore not
varied further. At each likely temperature, a number of models was calculated with different elemental abundances, as to ensure that
different abundance patterns would not affect the constraining of temperature.
For all objects
studied here, we found that a microturbulence velocity higher than that adopted
in the grid models (Table \ref{tab:bestfitmodels}) improves the fits to the O\,{\sc vi} $\lambda \lambda$ 1031.9,
1037.6 $\mathrm{\AA}$ P-Cygni profile, as shown in Fig. \ref{vturbNGC5189}, and
also to the C\,{\sc iv} $\lambda \lambda$ $1548.2$, $1550.8$ $\mathrm{\AA}$ line.
\begin{figure}
\centering
\includegraphics[scale=0.435]{ngc5189vturb0_newnorm.pdf}
\caption{CSPN NGC 5189 observed O\,{\sc vi} $\lambda \lambda$ $1031.9$, $1037.6$
$\mathrm{\AA}$ line (black continuous line) is compared with a grid model with
$T_{\ast}=165$ kK, $R_{t}=9.9R_{\odot}$,
$v_{\infty}=2500$
km s$^{-1}$ recomputed to account for different microturbulence
velocities.}\label{vturbNGC5189}
\vspace{0.5cm}
\end{figure}
\subsection{Results for individual objects}
The observed spectra and final best-fitting models for the individual objects
are shown in Figs. \ref{allfits_FUV}, \ref{allfits_UV}, and \ref{allfits_NUV}.
The parameters of our best-fitting models are given in Table
\ref{tab:bestfitmodels}. The uncertainties in stellar temperature are given in
the sections on the individual objects. Whenever far-UV, UV, and near-UV spectra
are available for an object, we estimate the uncertainty in $T_{\ast}$ to be
roughly half the interval between the values adopted in our grid of synthetic
spectra. When, as is the case of NGC 2867 and Sand 3, not all the necessary line
diagnostics are available, the uncertainties in temperature are deemed larger.
The uncertainties in the mass-loss rate are dominated by the uncertainties in
the distance and in the clumping filling factor, which was assumed a fixed
parameter throughout this work. We can typically constrain $R_{t}$ within $\pm$
2.0 R$_{\odot}$. Assuming the stellar parameters of the best-fitting models of
the various sample objects, we estimate the uncertainties of the carbon and
helium mass fraction are about $\pm0.10$, as illustrated in Fig.
\ref{carbonuncertainty}, and the uncertainties of the oxygen and nitrogen mass
fractions to be about $\pm0.06$ (Fig. 17 from \citet{kelleretal2011PORT}
illustrates the impact of different oxygen abundances on the synthetic spectra)
and $\pm0.01$ (Fig. \ref{nitrogenuncertainty}), respectively. We have used the
far-UV Ne\,{\sc vii} and Ne\,{\sc vi} features in the FUSE spectra to determine the
neon mass fractions. However, as we discuss further below, Ne features appear in
the synthetic spectra at longer wavelengths, which are not observed, perhaps due to uncertainties in the atomic data. The Ne\,{\sc vii} $\lambda$ $973.3$
$\mathrm{\AA}$ line, however, has been used to infer enhanced neon abundances
in PG 1159 stars by \citet{2005ApJ...627..424H} and \citet{2004A&A...427..685W}.
The atomic data used for the $\lambda$ $973.3$ $\mathrm{\AA}$ Ne\,{\sc vii} line are
described in \citet{2005ApJ...627..424H}.
The observed spectra shown in Figs. \ref{allfits_UV} and \ref{allfits_NUV} were
obtained with different instruments and thus, differ in spectral resolution.
Sand 3, NGC 6905, and Pb 6 STIS spectra were obtained with different apertures
and thus display different line spread functions (LSFs) as well. The synthetic
spectra were convolved with the corresponding LSFs when being compared to STIS
observations and with Gaussians of appropriate FWHM when compared to IUE and
FUSE data.
\subsubsection{CSPN NGC 6905 and Sand 3}
CSPN NGC 6905 and Sand 3 are the two stars for which we inferred lower
temperatures: $T_{\ast}=150\pm^{8}_{13}$ kK for NGC 6905 and
$T_{\ast}=150\pm^{8}_{25}$ kK for Sand 3, the difference in error bars being due
to the lack of FUSE spectra available for Sand 3, which deprives the analysis of
the far-UV O\,{\sc vi} and Ne\,{\sc vii} P-Cygni profiles, which are the line diagnostics that
most clearly separate models of 125,000 and 150,000 K. We determined terminal
velocities of the wind of $v_{\infty}=$2000$\pm 200$ and 2200$\pm 200$ km
s$^{-1}$ respectively. Their STIS spectra are very similar and both stars show O\,{\sc v} lines much stronger than in all other sample objects. Two of
the most conspicuous differences between the spectra of these two stars are the
strong N\,{\sc v} $\lambda \lambda$ 1238.8, 1242.8 $\mathrm{\AA}$ P-Cygni profile seen
in Sand 3's spectrum and absent from NGC 6905's and the even stronger O\,{\sc v}
$\lambda \lambda$ $2781.0$, $2787.0$ $\mathrm{\AA}$ line observed in Sand 3.
From the absence of a N\,{\sc v} $\lambda \lambda$ 1238.8, 1242.8 $\mathrm{\AA}$ line
profile in the observed spectra of NGC 6905, we derived, in
\citet{kelleretal2011PORT}, an upper limit to its N mass fraction of
$5.5\times10^{-4}$, while for Sand 3, a much higher abundance of $X_{N}=0.07$
was found. For both stars we found the same oxygen mass fraction,
$X_{O}=0.08\pm0.06$. Although neither best-fitting model for either star
reproduces the observed intensity of the O\,{\sc v} $\lambda \lambda$ $2781.0$,
$2787.0$ $\mathrm{\AA}$ line, our best-fitting model for Sand 3 predicts a
stronger line than our model for CSPN NGC 6905, which is due to the lower value
of transformed radius adopted in the former model ($R_{t}=9.4\pm2.0$ R$_{\odot}$
for Sand 3 and $R_{t}=10.7\pm2.0$ R$_{\odot}$ for NGC 6905). Because no FUSE
spectra are available for Sand 3, we simply kept the Ne abundance used in the
grid models.
Carbon and helium mass fractions derived by us for CSPN NGC 6905 ($X_{C}=0.45
\pm 0.10$, $X_{He}=0.45 \pm 0.10$) and Sand 3 ($X_{C}=0.55 \pm 0.10$, $X_{He}=0.28
\pm 0.10$) are quite different, which illustrates the fact that the overall
appearance of [WCE] spectra is mainly governed by temperature and mass loss, and
He and C abundances play only a secondary role, with big changes in
abundance producing minor changes in line strength.
Optical and low resolution IUE spectra of Sand 3 were previously analysed by
\cite{1997A&A...320...91K},
using non-LTE, homogeneous, spherically symmetric expanding atmosphere models,
which included helium, carbon,
oxygen and nitrogen in the calculations. They found $v_{\infty}=2200$ km/s,
$T_{\ast}=140000$ K, and $R_{t}=3.0$ R$_{\odot}$. As for the elemental mass
fractions, they derived $X_{He}=0.615$, $X_{C}=0.26$, $X_{O}=0.12$ and
$X_{N}=0.005$. Values of $v_{\infty}$, $T_{\ast}$, and $X_{O}$ are in agreement
with those derived in this work within the error bars. $X_{He}$, $X_{C}$, and
$X_{N}$ are in strong disagreement. In a previous work,
\citet{Barlow1982} derived He, C, and O mass fractions that are close to the
values derived by us. A previous analysis of CSPNe NGC 6905 was discussed in
\citet{kelleretal2011PORT}.
\subsubsection{CSPN NGC 2867}
The O\,{\sc vi} $\lambda \lambda$ 1031.9, 1037.6 $\mathrm{\AA}$ P-Cygni profile of CSPN
NGC 2867 is clearly narrower than those observed in NGC 5189 and Pb 6 spectra
and thus we derived $v_{\infty}=$2000$\pm 250$ km s$^{-1}$ for the central star
of NGC 2867. If indeed the nebula-dominated low resolution IUE spectra of NGC
2867 contains the stellar flux, we can set an upper limit to the strength of the O\,{\sc v}
lines and thus better constrain the temperature. Since no strong O\,{\sc v} lines are
seen, models with stellar temperatures of 165,000 K would be favoured. If not,
we can only leverage on FUSE spectra to discern between model temperatures.
However, this spectral region does not show remarkable differences between
models of 150,000 and 165,000 K, although it rules out models of $T_{\ast} \le$
125,000 K which do not produce the strong Ne\,{\sc vii} and O\,{\sc vi} P-Cygni profiles
observed. Models of 165,000 K do produce, however, somewhat stronger Ne\,{\sc vii} and
O\,{\sc vi} lines, while concomitantly producing reasonable fits for the other
available lines, than models of 150,000 K, in which these lines appear somewhat
weaker than observed. We thus derived $T_{\ast}=165\pm^{18}_{20}$ kK for the
central star of NGC 2867, $R_{t}=8.5\pm2.0$ R$_{\odot}$, $X_{C}=0.25 \pm0.10$,
$X_{He}=0.60 \pm 0.10$ (even though we did not attempt a fit of the He\,{\sc ii} 1640.4
$\mathrm{\AA}$ line due to the heavy nebular contamination in the observed
spectra of this star, its helium mass fraction can be determined for it is the
complement of the sum of all other elemental mass fractions considered in the
calculations), and $X_{O}=0.10 \pm 0.06$. Since no nitrogen line is available in
the FUSE range, we merely kept the nitrogen mass fraction adopted in our grid
models.
\citet{1997IAUS..180..114K} analysed optical and IUE spectra of a series of
CSPNe,
among them the central stars of NGC 2867, NGC 5189, and Pb 6 using a non-LTE
radiative transfer code
that considers homogeneous winds and He, C, O, and N opacities. They found, for
NGC 2867, $v_{\infty}=1800$ km s$^{-1}$, $T_{\ast}=141000$ K, and $R_{t}=4.0$
R$_{\odot}$. The elemental mass fractions derived by them are $X_{He}=0.66$,
$X_{C}=0.25$, and $X_{O}=0.09$. Their derived values of $v_{\infty}$, $X_{He}$,
$X_{C}$, and $X_{O}$ are in agreement within the errors.
\subsubsection{CSPN Pb 6} \
CSPN Pb 6, as well as NGC 5189's central star, displays wider O\,{\sc vi} $\lambda
\lambda$ 1031.9, 1037.6 $\mathrm{\AA}$ and C\,{\sc iv} $\lambda \lambda$ $1548.2$,
$1550.8$ $\mathrm{\AA}$ P-Cygni profiles than the other sample objects. We
determined a terminal velocity of the wind of $v_{\infty}=$2500$\pm 200$ km
s$^{-1}$ for this star. The strong O\,{\sc vi} $\lambda \lambda$ 1031.9, 1037.6
$\mathrm{\AA}$ far-UV P-Cygni profile and weak O\,{\sc v} lines observed in CSPN Pb 6
suggest a temperature of $T_{\ast}=165\pm^{18}_{8}$ kK. The derived transformed
radius and elemental abundances are: $R_{t}=9.9\pm2.0$ R$_{\odot}$, $X_{C}=0.35
\pm 0.10$, $X_{He}=0.49 \pm 0.10$, $X_{O}=0.12 \pm 0.06$, $X_{N}=0.03 \pm
0.01$.
\citet{1997IAUS..180..114K} found for this star: $v_{\infty}=3000$ km s$^{-1}$,
$T_{\ast}=140000$ K, $R_{t}=4.5$ R$_{\odot}$, $X_{He}=0.617$, $X_{C}=0.24$,
$X_{O}=0.14$, and $X_{N}=0.003$. Only $X_{O}$ is in agreement with values derived
in this work within the errors. The stellar temperature derived by us is
considerably higher and the terminal velocity is much lower. Especially
discrepant is the nitrogen mass fraction, whose value derived by us is 10 times
that of the previous analysis.
\subsubsection{CSPN NGC 5189} \
\label{sec:5189}
As in the case of CSPN Pb 6, we found a terminal velocity of the wind of
$v_{\infty}=$2500$\pm 250$ km s$^{-1}$ for the central star of NGC 5189. The
presence of very strong far-UV Ne\,{\sc vii} and O\,{\sc vi} P-Cygni profiles in its FUSE
spectrum, accompanied by weak O\,{\sc v} lines, led us to derive a stellar temperature
of $T_{\ast}=165\pm^{18}_{8}$ kK, similarly to Pb 6 and NGC 2867. We derived,
for the central star of NGC 5189, $R_{t}=9.9\pm2.0$ R$_{\odot}$ (same as for Pb
6), $X_{C}=0.25 \pm 0.10$ (same as for NGC 2867), $X_{He}=0.58 \pm 0.10$,
$X_{O}=0.12 \pm 0.06$ (same as for Pb 6), $X_{N}=0.01 \pm 0.01$.
Although we regard the He\,{\sc ii} $\lambda$ 1640.4 $\mathrm{\AA}$ line as a mere upper limit to
the stellar contribution, due to the evidence of nebular contamination presented
above, and despite the fact that we derived the parameters of our best-fitting
model without attempting to precisely match the observed intensity of this line,
it is, however, reproduced in the final model. In order to maintain the overall
fit to the other wind lines and at the same time decrease the strength of the He\,{\sc ii} 1640.4 $\mathrm{\AA}$ line, the helium mass fraction would have to be
severely lowered, at the same time that the value of mass-loss rate would have
to be decreased as to allow the correspondent increase in carbon and oxygen mass
fraction that need to occur to assimilate the decrease in the helium mass
fraction. Mass fractions of neon and nitrogen would also have to be revised up
to compensate for the lower mass-loss rate. We computed models to explore this
alternative. They result in considerably worse fits to the O\,{\sc vi} $\lambda
\lambda$ 1031.9, 1037.6 $\mathrm{\AA}$ doublet, which is now very weak in
comparison to the observations. This incongruence illustrates the need for higher resolution UV
spectra of CSPNe.
The stellar parameters derived by \citet{1997IAUS..180..114K} for NGC 5189 are
$v_{\infty}=3000$ km/s, $T_{\ast}=135000$ K, $R_{t}=5.0$ R$_{\odot}$, and
$X_{He}=0.757$, $X_{C}=0.16$, $X_{O}=0.08$, and $X_{N}=0.003$, in mass fractions.
Only carbon and oxygen mass fractions derived by them agree within the errors
with those found by us. As in the case of Pb 6, a much higher stellar
temperature and considerably lower terminal velocity of the wind and helium mass
fraction were derived by us, besides a 10 times higher nitrogen mass fraction.
\begin{table}
\caption{\small Parameters of our best-fitting models for the sample objects and
distance dependent parameters. $X_{He}$, $X_{C}$, $X_{N}$, $X_{O}$, and $X_{Ne}$
are the
helium, carbon, nitrogen, oxygen, and neon mass fractions, respectively. For the
discussion on the derived parameters and errors, see text. Results by other
authors are also shown. }
\scriptsize
\tabcolsep=0.09cm
\begin{tabular}{ccccccccccccc}
\hline
Object & Pb 6 & & NGC 6905 & &
& NGC 5189 & & Sand 3 & & & NGC 2867
& \\
\hline
&t.w. & KH1997a & t.w. & M2007 &
KH1997a & t.w. & KH1997a & t.w. & KH1997b & BH1982 & t.w.
& KH1997a\\
\hline
$T_{\ast}$ [kK] & 165 & 140 & 150 & 149.6 & 141
& 165 & 135 & 150 & 140 & 200 & 165
& 141 \\
$R_{t}$ [R$_{\odot}$] & 9.9 & 4.5 & 10.7 & 10.5 & 3.4
& 9.9 & 5.0 & 9.4 & 3.0 & & 8.5
& 4.0 \\
$v_{\infty}$ [km/s] & 2500 & 3000 & 2000 & 1890 & 1800
& 2500 & 3000 & 2200 & 2200 & 2700 & 2000
& 1800 \\
$X_{He}$ & 0.49 & 0.617 & 0.45 & 0.49 & 0.60
& 0.58 & 0.757 & 0.28 & 0.615 & 0.38 & 0.60
& 0.66 \\
$X_{C}$ & 0.35 & 0.24 & 0.45 & 0.40 & 0.25
& 0.25 & 0.16 & 0.55 & 0.26 & 0.54 & 0.25
& 0.25 \\
$X_{O}$ & 0.12 & 0.14 & 0.08 & 0.10 & 0.15
& 0.12 & 0.08 & 0.08 & 0.12 & 0.08 & 0.10
& 0.09 \\
$X_{N}$ & 0.03 & 0.003 & 0.00011 &$<$0.001& 0
& 0.01 & 0.003 & 0.07 & 0.005 & &
0.01\tablenotemark{b} & 0 \\
$X_{Ne}$ & 0.01 & & 0.02 & &
& 0.04 & & 0.02\tablenotemark{b}&& & 0.04
& \\
$\xi_{max}$ [km/s] & 200 & & 150 & &
& 200 & & 150 & & & 200
& \\
d\tablenotemark{c} [kpc] & 4.38 & & 1.70 & &
& 0.55 & & 0.80\tablenotemark{a}&& & 1.84
& \\
\hline
distance dependent &&&&&&&&&&&& \\
parameters &&&&&&&&&&&& \\
R$_{\ast}$ [R$_{\odot}$] & 0.06 & & 0.09 & &
& 0.03 & & 0.12\tablenotemark{a} & & & 0.05
&\\
$\log L/L_{\odot}$ & 3.43 & & 3.56 & &
& 2.73 & & 3.84\tablenotemark{a} & & &3.15
& \\
$log \dot{M}$ [M$_{\odot}$yr$^{-1}$] & -7.29 & & -7.21 & &
& -7.81 & & -6.89\tablenotemark{a}& & & -7.50
& \\
\hline
\label{tab:bestfitmodels}
\end{tabular}
\small{t.w.=this work; KH1997a=\citet{1997IAUS..180..114K};
KH1997b=\citet{1997A&A...320...91K}; M2007=\citet{2007ApJ...654.1068M},
BH1982=\citet{Barlow1982}}
\vspace{-0.2in}
\tablenotetext{a}{Values obtained assuming $\log L/L_{\odot}$=3.84.}
\tablenotetext{b}{Not constrained. Value assumed on the grid models was kept.}
\tablenotetext{c}{Distances from Table \ref{objects}.}
\end{table}
\vspace{0.1in}
\subsection{The problematic near-UV region}
In the studied objects, the spectral region between 1700 and 3200 $\mathrm{\AA}$
is problematic.
The available spectra have worse resolution in this region, which is full of
structures that make the
determination of the stellar continuum uncertain. In the region between 1700 and
2400 $\mathrm{\AA}$, observed features both in absorption and
in emission are not reproduced by our best-fitting models or grid models of any
temperature and mass-loss rate. The region between 2150 and 2300 $\mathrm{\AA}$
is particularly challenging.
There is, in this region, a strong feature observed in the central stars of NGC
6905, NGC 5189, Pb 6, and Sand 3, showing an emission
component on the red side and an absorption one on the blue side. This feature
reminds us of a P-Cygni profile in the
observed spectra of Pb 6 and of NGC 6905 central stars,
but in the noisier and low resolution spectra available for NGC 5189 and Sand 3,
the blue side resembles a simple absorption. Our models are able to reproduce
the red side of these features, in the case of NGC 6905 and of NGC 5189, but not
the blue absorptions that are located very close to the broad feature of the
interstellar extinction curve at 2175 $\mathrm{\AA}$, which is related to
graphite grains \citep{Draine1993}.
Another difficulty in modelling these stars consists of weak Ne\,{\sc v} and Ne\,{\sc vi}
lines that are predicted throughout the near-UV by the models,
but are not observed. While we checked CMFGEN's neon atomic data files for
consistency with the Atomic Line List
(\url{http://www.pa.uky.edu/~peter/atomic/}) and found no discrepancy, there may be uncertainties in the atomic data. The neon
atomic data are primarily taken from the Opacity Project
\citep{1987JPhB...20.6363S,1994MNRAS.266..805S} and the Atomic Spectra Database
at the NIST Physical Measurement Laboratory.
\subsection{Argon and Iron}
In \citet{kelleretal2011PORT}, we found that a model with 10 times the argon
solar abundance produces an Ar\,{\sc vii} $\lambda$ 1063.55 $\mathrm{\AA}$ line which
approximately reproduces the observed line profile in that spectral region of
CSPN NGC 6905. In Fig. \ref{allfits_FUV}, the best-fitting models for NGC 5189
and NGC 2867 also have the same argon overabundance. However, the shapes of
the observed and synthetic profiles disagree, challenging the identification of
this line.
In all best-fitting models, we assumed a strong iron underabundance of one
hundredth of the solar value, since higher values would entail the appearance of a number of non-observed Fe\,{\sc x} lines in the synthetic spectra
\citep{kelleretal2011PORT}. These lines are not, however, observed in PG1159
stars for which solar iron abundances were found by \citet{Werner2010,Werner2011a} and there may be strong uncertainties in their computed
wavelengths.
\subsection{Reddening}
Having constrained the main stellar parameters through the analysis of the
strength of spectral lines, we now compare the slope of the observed
spectral continuum of each sample object to that of our best-fitting models
reddened by different values of colour excess as shown in Fig. \ref{Reddening}.
We adopted \citet{Cardelli1989} extinction curves with $R_{V}=3.1$. The results
are listed in Table \ref{tab:is}. In the case of NGC 2867, which is shown
separately in Fig. \ref{NGC2867Reddening}, we had to take into consideration the
contribution from the nebular continuum, which heavily affects
wavelengths longer than Ly\,$\alpha$. We estimated the nebular continuum
emission using the code described in \citet{1997ApJ...480..290B}, which accounts
for two-photon, H and He recombination, and free-free emission processes.
Nebular parameters were taken from the literature
\citep{2002ApJS..138..285M,2002ApJS..138..279M} and they include the electron
density, electron temperature, observed H$\beta$ flux, helium-to-hydrogen ratio,
ratio of doubly to singly ionized helium and the extinction quantity ($c$).
We found $E(B-V)=0.18$, which corresponds to the value of
$c$ found in the literature, to allow for an overall good match between observed
continuum of NGC 2867 and the sum of model stellar and nebular emissions.
\begin{figure}
\begin{center}
\includegraphics[scale=0.42]{paperplot_wce__newvelnewmasslossSand3_neb0_final.pdf}
\caption{FUSE spectra (black line) of CSPNe Pb 6, NGC 5189, NGC 2867, and NGC
6905 shown together with our best-fitting models, with (pink) and without (blue)
the effects of interstellar absorption due to molecular and atomic hydrogen. The
observed and synthetic spectra were degraded to a 0.5 $\mathrm{\AA}$ resolution
for clarity. No FUSE is spectrum is available for Sand 3. We, however, still
show the synthetic spectra prediction for this wavelength interval with a 0.5
$\mathrm{\AA}$ resolution for comparison with results for the other sample
objects. HST/STIS spectra of this star comprehend the C\,{\sc iv} $\lambda\lambda$
1168.8, 1169.0 $\mathrm{\AA}$ line, which is shown here in black. In pink and
green we show our best-fitting model for Sand 3, with and without Ly\,$\alpha$
absorption, convolved with the appropriate HST/STIS LSF. The identifications of
stellar lines shown in this and following figures correspond to the strongest
lines predicted by the models at each spectral region.}
\label{allfits_FUV}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[scale=0.435]{paperplot_wce__newvelnewmasslossSand3_neb1_final.pdf}
\caption{Observed spectra of the sample objects (black) in the region between
1200 and 1750 $\mathrm{\AA}$ shown with our best-fitting models, with (pink) and
without (blue) ISM absorption, degraded to match the resolution of the
observations (see table \ref{spectra}). For CSPN NGC 5189 we also show the
subtracted spectrum from Fig. \ref{subtraction}, in yellow. The grey labels on
the top part of the figure identify the position of commonly present nebular
lines.}
\label{allfits_UV}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[scale=0.435]{paperplot_wce__newvelnewmasslossSand3_neb2_final.pdf}
\caption{Near-UV observed spectra (black) together with our best-fitting models
to the sample objects (blue).}
\label{allfits_NUV}
\end{center}
\end{figure}
\begin{figure}
\centering
\includegraphics[angle=0,scale=0.5]{ngc6905_t150_C_1_norm.pdf}
\caption{Observed spectrum of the central star of NGC 6905 (black continuous
line) and models with $T_{\ast}=150$ kK,
$R_{t}=10.7R_{\odot}$, $v_{\infty}=2000$ km s$^{-1}$ and different carbon and
helium abundances.} \label{carbonuncertainty}
\end{figure}
\begin{figure}
\centering
\includegraphics[angle=0,scale=0.5]{NGC5189_t165_N165_0_newnorm.pdf}
\caption{Observed spectra of the central star of NGC 5189 (black continuous
line) and models with $T_{\ast}=165$ kK,
$R_{t}=9.9R_{\odot}$, $v_{\infty}=2500$ km s$^{-1}$ and different nitrogen
abundances.} \label{nitrogenuncertainty}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth,height=0.185\textheight]{PNPb6_final_ebv_310000test_final.pdf}
\includegraphics[width=0.9\linewidth,height=0.185\textheight]{NGC5189_final_ebv_3_named_final.pdf}
\includegraphics[width=0.9\linewidth,height=0.185\textheight]{NGC6905_final_ebv_3_named_final.pdf}
\includegraphics[width=0.9\linewidth,height=0.185\textheight]{Sand3_final_ebv_Rv31_final.pdf}
\caption{The observed spectra of the sample objects (continuous black lines),
except for NGC 2867,
are shown together with our best-fitting models reddened by different values of
the colour excess.
We used the \citet{Cardelli1989} extinction curves, with $R_{V}=3.1$.}
\label{Reddening}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth,height=0.185\textheight]{ngc2867_nebularcontinuum_EBV010_final.pdf}
\includegraphics[width=0.9\linewidth,height=0.185\textheight]{ngc2867_nebularcontinuum_EBV018_final.pdf}
\includegraphics[width=0.9\linewidth,height=0.185\textheight]{ngc2867_nebularcontinuum_EBV024_final.pdf}
\caption{The observed spectra of NGC 2867 (continuous black lines) dereddened
adopting different values of the colour excess is shown together with our
best-fitting model for the stellar spectrum (green dotted line), a model for the
nebular continuum (blue continuous line) and the sum of both models (cyan dashed
line).
We used the \citet{Cardelli1989} extinction curves, with $R_{V}=3.1$.}
\label{NGC2867Reddening}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[scale=0.435]{TxL_GRID_paperII_final.pdf}
\caption{HR diagram with evolutionary tracks from \citet{2006A&A...454..845M}.
The open circles correspond to the sample objects. Sand 3 was omitted from it,
because, to the best of our knowledge, there is no measurement of its distance
in the literature.}
\label{TxL_GRID}
\end{center}
\end{figure}
\subsection{Discussion}
\label{subsec:abund}
Fig. \ref{TxL_GRID} shows the evolutionary tracks of \citet{2006A&A...454..845M}
in the $\log(T_{eff}) \times \log(L)$ diagram and the position of our sample
objects. Luminosities are based on distances taken from the literature and are
shown in Table \ref{tab:bestfitmodels}. Sand 3 was omitted from the figure,
because we were unable to find measurements of its distance in the literature.
The sample objects intercept the evolutionary tracks at 0.565 and 0.584 solar
masses. The distance values adopted from the literature place the central stars
of NGC 6905 and Pb 6 on the verge of entering the WD cooling track, while they
put the central stars of NGC 5189 and NGC 2867 already within it, in a region
where PG1159 stars are expected. [WCE]-type CSPNe are, however, expected to
populate the region of constant luminosity of the evolutionary tracks in the
HRD, which, at the temperatures derived here, would require higher luminosities
and thus, greater distances.
Among the five objects analysed in this work, Sand 3 is the only one whose
derived abundances
seem to place it outside the range of values found in the literature for [WC]
stars, since it shows,
according to Table \ref{tab:bestfitmodels}, a quite low helium abundance of 0.28
in mass fraction (not significantly lower, however, than the lowest value of
helium mass fraction found for PG1159 objects, which is 0.30).
The derived He, C, and O abundances, in mass fractions, of our sample objects are
in the intervals: $X_{He}=0.28-0.60$, $X_{C}=0.25-0.55$, and $X_{O}=0.08-0.12$
(see Fig. \ref{abunds}). It is generally believed that [WCL]-type CSPNe
evolve into [WCE] objects and thus, both classes are expected to present
somewhat similar carbon to helium ratios. However, previous analyses have shown
the carbon content in [WCE]-type stars to be lower than what is seen in [WCL]
objects, with 0.19$<C:He<$0.7 for [WCE] stars
\citep{1997IAUS..180..114K,1997A&A...320...91K} and $C:He$ typically higher than
1 for [WCL] stars \citep{1996A&A...312..167L}. The wide range of values found by
us in this work seems to argue against this notion: we found
$C:He=$0.42, 0.43, 1.0, 0.71, 1.96 for the central stars of NGC 2867, NGC 5189,
NGC 6905, Pb 6, and Sand 3, respectively. This is in line with the results from
\citet{2003IAUS..209..243C} and \citet{2007ApJ...654.1068M} that found no
systematic discrepancies in the $C:He$ ratios between [WCL] and [WCE] stars.
Values of the nitrogen mass fraction derived in the literature range from
$< 3 \times 10^{-5}$ \citep[see][and references therein]{2006PASP..118..183W}
to 0.04 \citep{Leuenhagen1993}. We have derived an upper limit of
$5.5\times10^{-4}$ for the nitrogen abundance of NGC 6905, which would indicate
it is the product of a LTP or even an AFTP. The central stars of NGC 5189, Pb 6
and Sand 3, on the other hand, all show conspicuous N\,{\sc v} $\lambda \lambda$
1238.8, 1242.8 $\mathrm{\AA}$ lines that led to derived
nitrogen mass fractions of 1, 3, and 7 per cent, respectively, pointing to the
occurrence of VLTP events.
We have not included hydrogen in our grid models or in the best-fitting models
presented in this work. However, we calculated exploratory models which indicate
an upper limit to the hydrogen mass fraction of 20 per cent for our sample
objects.
Neon abundances of about 2 per cent have been reported for a number of PG1159
stars \citep{2004A&A...427..685W}, which is in agreement with stellar evolution
models that predict a neon mass fraction of about 2 per cent in the intershell
region. In this work, modelling of the far-UV Ne\,{\sc vii} and Ne\,{\sc vi} features led us
to infer neon mass fractions between 1 and 4 per cent.
\begin{figure}
\begin{center}
\includegraphics[scale=0.435]{abundancespaper2.pdf}
\caption{The abundance intervals of PG1159, [WC]-PG1159 and [WC] stars (as found
from the literature) are compared to the values determined in this work.}
\label{abunds}
\end{center}
\end{figure}
\section{Conclusions}
\label{sec:conclusions}
We have analysed UV spectra from 5 of the hottest known [WCE]-type CSPNe, which are
the central stars of NGC 2867, NGC 5189, NGC 6905, Pb 6, and Sand 3. The
analysis was performed using our
extensive grid of synthetic spectra, calculated with CMFGEN, which allowed us
to constrain temperatures, transformed radii, and terminal velocities of the
wind, in a uniform and systematic way. The
analysis was then refined by exploring other parameters, such as elemental abundances and microturbulence velocities, besides adding
further elements and ionic species into the calculations.
We applied the effects of interstellar absorption due to neutral and
molecular hydrogen to our best-fitting
models and determined their column densities and after the stellar parameters
were constrained based on line diagnostics, we compared the
observed slope of the stellar continua to our best-fitting models reddened using
different values
of colour excess.
We derived $T_{\ast} = 150000$ K for CSPNe NGC 6905 and Sand 3 and $T_{\ast} =
165000$ K for the central stars of NGC 2867, NGC 5189, and Pb 6. This result
corresponds to an upward revision of the temperatures derived for the latter
three objects, in comparison with previous analyses (for Sand 3 and the
central star of NGC 6905, the derived temperatures agree with previous results
within the errors), a fact that has two reasons: the first one is the inclusion
of several less abundant ions into the calculations, as described above, that
caused the O\,{\sc v} lines to increase in intensity without altering other line
diagnostics, allowing us to fit the observed spectra with hotter models. The O\,{\sc v}
lines are important temperature diagnostics and allowed us to separate between
models of 150 and 165 kK; The second reason is the modelling of the O\,{\sc vi} $\lambda
\lambda$ 1031.9, 1037.6 $\mathrm{\AA}$ and Ne\,{\sc vii} $\lambda$ 973 $\mathrm{\AA}$
lines in the FUSE range, which had not been modelled in previous analyses of
these three stars and which are stronger in our grid models of 165 kK. The
terminal wind velocities of the sample objects were found to range from 2000 to
2500 km s$^{-1}$. Despite the fact that the five stars studied here present
somewhat similar spectra, and that
we derived similar temperatures for them, we derived quite different helium,
carbon
and nitrogen abundances. The values of carbon to helium mass ratio found by us
for the sample objects span a wide range of values: $0.42<C:He<1.96$. The case
of Sand 3 seems an intriguing
one. Not only we find a very low helium abundance for this star, but we also
derive a high
nitrogen content, when comparing these with values found in the literature for
[WC] and PG1159
stars. Nitrogen abundances derived by us for the central stars of NGC 5189, Pb 6
and Sand 3 were found to be higher than previous analyses by factors of 3, 10
and 14, respectively.
By considering many elements and ionic species neglected in previous analyses, we improved the fits to the observed spectra: we were able to
reproduce the observed
strengths of the O\,{\sc v} $\lambda$ 1371.3 $\mathrm{\AA}$ lines and improve the fits
of the
O\,{\sc v} $\lambda \lambda$ 2781.0, 2787.0 $\mathrm{\AA}$ lines,
thanks to the consequent increase of the O\,{\sc v} ionization fractions and we were
also provided with the Ne\,{\sc vii} $\lambda$ 973 $\mathrm{\AA}$ line diagnostic,
which was modelled for the first time in [WCE] stars. The modelling of the
far-UV Ne features points towards strong neon overabundances for the central
stars of NGC 6905, NGC 5189, NGC 2867, and Pb 6, with Ne mass fractions between
0.01 and 0.04.
The near-UV region remains problematic: observed features both in absorption
and
in emission are not reproduced by our best-fitting models or grid models of any
temperature and mass-loss rate and weak Ne\,{\sc v} and Ne\,{\sc vi} lines are predicted by
the models throughout the region, but are not observed.
\vspace{1cm}
\textit{Acknowledgements.} We thank Klaus Werner for providing many useful
comments. We are also grateful to an anonymous referee for several comments
that helped us improve the present paper. G.R. Keller gratefully acknowledges financial support from the
Brazilian agencies CAPES (which supported her work at the Johns Hopkins
University, Department of Physics and Astronomy) and FAPESP,
grants 0370-09-6, 06/58240-3, and 2012/03479-2. This work was based on
observations made with the Far Ultraviolet Spectroscopic Explorer,
the Hubble Space Telescope, and the International Ultraviolet Explorer. The
data presented
in this paper were obtained from the Mikulski Archive for Space Telescopes
(MAST). STScI is operated by the Association of Universities for Research in
Astronomy, Inc., under NASA contract NAS5-26555. Support for MAST for non-HST
data is provided by the NASA Office of Space Science via grant NNX13AC07G and by
other grants and contracts. This work has made use of the computing facilities
of the Laboratory of Astroinformatics (IAG/USP, NAT/Unicsul), whose purchase was
made possible by the Brazilian agency FAPESP (grant 2009/54006-4) and the
INCT-A.
\bibliographystyle{grk_pp_mnras}
|
\section{The assumptions}
The assumptions in this paper---and in most theoretical papers on high-dimensional regression---have several components.
These include:
\begin{longlist}[(5)]
\item[(1)] The linear model is correct.
\item[(2)] The variance is constant.
\item[(3)] The errors have a Normal distribution.
\item[(4)] The parameter vector is sparse.
\item[(5)] The design matrix has very weak collinearity.
This is usually stated in the form of
incoherence, eigenvalue restrictions or incompatibility assumptions.
\end{longlist}
To the best of my knowledge, these assumptions are not testable when $p
> n$.
They are certainly a good starting place for
theoretical investigations but they are
indeed very strong.
The regression function
$m(x) = \mathbb{E}(Y|X=x)$ can be any function.
There is no reason to think it will be close to linear.
Design assumptions are also highly suspect.
High collinearity is the rule rather than the exception especially
in high-dimensional problems.
An exception is signal processing,
in particular compressed sensing,
where the user gets to construct the design matrix.
In this case, if the design matrix is filled with
independent random Normals, the design matrix will
be incoherent with high probability.
But this is a rather special situation.
None of this is meant as a criticism of the paper.
Rather, I am trying to motivate interest in the question
I asked earlier, namely:
what can we do without these assumptions?
\begin{remark}
It is also worth mentioning that even in
low-dimensional models and even if the model is correct,
model selection raises troubling issues that
we all tend to ignore.
In particular, variable selection
makes the minimax risk explode
[\citeauthor{LeePot08} (\citeyear{LeePot05,LeePot08})].
This is not some sort of pathological risk explosion, rather,
the risk is large in a neighborhood of 0,
which is a part of the parameter space we care about.
\end{remark}
\section{The assumption-free lasso}
To begin it is worth pointing out that the lasso
has a very nice assumption-free interpretation.
Suppose we observe
$(X_1,Y_1),\ldots, (X_n,Y_n)\sim P$
where
$Y_i\in\mathbb{R}$ and
$X_i \in\mathbb{R}^d$.
The regression function
$m(x) = \mathbb{E}(Y|X=x)$
is some unknown, arbitrary function.
We have no hope to estimate $m(x)$
nor do we have licence to impose assumptions on $m$.
But there is sound theory to justify the lasso
that makes virtually no assumptions.
In particular, I refer to
\citet{GreRit04} and
\citet{JudNem00}.
Let
${\mathcal L} = \{ x^t\beta\dvtx \beta\in\mathbb{R}^d\}$
be the set of linear predictors.
For a given $\beta$,
define the predictive risk
\[
R(\beta) = E\bigl( Y - \beta^T X\bigr)^2,
\]
where $(X,Y)$ is a new pair.
Let us define the best, sparse, linear predictor
$\ell_*(x) = \beta_*^T x$
(in the $\ell_1$ sense)
where
$\beta_*$ minimizes
$R(\beta)$ over the set
$B(L) = \{ \beta\dvtx \|\beta\|_1 \leq L\}$.
The lasso estimator $\hat\beta$
minimizes the empirical risk
$\hat R(\beta) = \frac{1}{n}\sum_{i=1}^n (Y_i - \beta^T X_i)^2$
over $B(L)$.
For simplicity, I will assume that
all the variables are bounded by $C$
(but this is not really needed).
We make no other assumptions:
no linearity,
no design assumptions
and no models.
It is now easily shown that
\[
R(\hat\beta) \leq R(\beta_*) + \sqrt{ \frac{8 C^2 L^4}{n} \log\biggl(
\frac{2p^2}{\delta} \biggr)}
\]
except on a set of probability at most $\delta$.
This shows that the predictive risk of the lasso
comes close to the risk of the best sparse linear predictor.
In my opinion, this explains why the lasso ``works.''
The lasso gives us a predictor with
a desirable property---sparsity---while being computationally tractable
and it comes close to the risk of the
best sparse linear predictor.
\section{Interlude: Weak versus strong modeling}
When developing new meth\-odology, I think it is useful
to consider three different stages of development:
\begin{longlist}[(3)]
\item[(1)] Constructing the method.
\item[(2)] Interpreting the output of the method.
\item[(3)] Studying the properties of the method.
\end{longlist}
I also think it is useful to
distinguish two types of modeling.
In \textit{strong modeling}, the model
is assumed to be true in all three stages.
In \textit{weak modeling}, the model
is assumed to be true for stage 1 but not for
stages 2 and 3.
In other words, one can use a model
to help construct a method.
But one does not have to assume the model is true
when it comes to interpretation or when
studying the theoretical properties of the method.
My discussion is guided by my preference
for weak modeling.
\section{Assumption-free inference: The HARNESS}
Here, I would like to discuss an approach I have been developing
with Ryan Tibshirani.
We call this:
{H}igh-dimensional Agnostic
{R}egression
{N}ot
{E}mploying
{S}tructure or
{S}parsity,
or, the HARNESS.
The method is a variant of the idea proposed in \citet{WasRoe09}.
The idea is to split the data into two halves.
${\mathcal D}_1$ and ${\mathcal D}_2$.
For simplicity, assume that $n$ is even so that each half has size
$m=n/2$.
From the first half ${\mathcal D}_1$, we select a subset of variables $S$.
The method is agnostic about how the variable selection is done.
It could be forward stepwise, lasso, elastic net or anything else.
The\vspace*{2pt} output of the first part of the analysis is
the subset of predictors
$S$ and an estimator
$\hat\beta= (\hat\beta_j\dvtx j\in S)$.
The second half of the data ${\mathcal D}_2$ is used to provide
distribution-free inferences for the following questions:
\begin{longlist}[(3)]
\item[(1)] What is the predictive risk of $\hat\beta$?
\item[(2)] How much does each variable in $S$ contribute to the predictive risk?
\item[(3)] What is the best linear predictor using the variables in $S$?
\end{longlist}
All the inferences from ${\mathcal D}_2$
are interpreted as being conditional on ${\mathcal D}_1$.
(A~variation is to
use ${\mathcal D}_1$ only to produce $S$
and then construct the coefficients of the predictor from
${\mathcal D}_2$.
For the purposes of this discussion, we use $\hat\beta$
from ${\mathcal D}_1$.)
In more detail,
let
\[
R = \mathbb{E}\bigl| Y-X^T \hat\beta\bigr|,
\]
where the randomness is over the new pair $(X,Y)$;
we are conditioning on ${\mathcal D}_1$.
Note that in this section I have changed
the definition of $R$ to be on the absolute scale
which is more interpretable.
In the above equation, it is understood
that \mbox{$\hat\beta_j=0$} when $j\notin S$.
The first question refers to producing a estimate and confidence
interval for $R$
(conditional on ${\mathcal D}_1$).
The second question refers to inferring
\[
R_j = \mathbb{E}\bigl| Y-X^T \hat\beta_{(j)}\bigr| -
\mathbb{E}\bigl| Y-X^T \hat\beta\bigr|\vadjust{\goodbreak}
\]
for each $j\in S$,
where
$\beta_{(j)}$ is equal to $\hat\beta$
except that $\hat\beta_j$ is set to 0.
Thus,
$R_j$ is the risk inflation by excluding $X_j$.
The third question refers to
inferring
\[
\beta^* = \argmin_{\beta\in\mathbb{R}^k} \mathbb{E}\bigl(Y - X_S^T
\beta\bigr)^2
\]
the coefficient of the best linear predictor for the chosen model.
We call $\beta^*$ the
\textit{projected parameter}.
Hence, $x^T\beta^*$ is
the best linear approximation to $m(x)$
on the linear space spanned by the selected variables.
A consistent estimate of $R$ is
\[
\hat{R} = \frac{1}{m}\sum_{i=1}^m
\delta_i,
\]
where the sum is over ${\mathcal D}_2$, and
$\delta_i = |Y_i - X_i^T \hat\beta|$.
An approximate $1-\alpha$ confidence interval for $R$ is
$\hat R \pm z_{\alpha/2} s/\sqrt{m}$
where
$s$ is the standard deviation of the $\delta_i$'s.
The validity of this confidence interval is essentially
distribution-free.
In fact, if we want to be purely distribution-free
and avoid asymptotics,
we could instead define $R$ to be the median of
the law of $|Y-X^T \hat\beta|$.
Then the order statistics of the $\delta_i$'s can be used
in the usual way to get a finite sample, distribution-free
confidence interval for $R$.
Estimates and confidence intervals for $R_j$ can be obtained from
$e_1,\ldots, e_m$
where
\[
e_i = \bigl|Y_i - X^T \hat\beta_{(j)}\bigr|
- \bigl|Y_i - X^T \hat\beta\bigr|.
\]
Estimates and confidence intervals for $\beta_*$
can be obtained by standard least squares procedures
based on ${\mathcal D}_2$.
The steps are summarized in Figure~\ref{figharness}.
\begin{figure}[b]
\hrulefill\vspace*{6pt}
\begin{center}
{\sf The HARNESS}\vspace*{6pt}
\end{center}
\flushleft{Input: data ${\mathcal D} = \{(X_1,Y_1),\ldots, (X_n,Y_n)\}$.}
\begin{longlist}[(3)]
\item[(1)] Randomly split the data into two halves
${\mathcal D}_1$ and ${\mathcal D}_2$.
\item[(2)] Use ${\mathcal D}_1$ to select a subset of variables $S$.
This can be forward stepwise, the lasso, or any other method.
\item[(3)] Let $R = \mathbb{E}( (Y-X^T \hat\beta)^2 | {\mathcal D}_1)$
be the predictive risk of the selected model
on a future pair $(X,Y)$, conditional on ${\mathcal D}_1$.
\item[(4)] Using ${\mathcal D}_2$ construct point estimates and confidence intervals
for $R$, $(R_j\dvtx \hat\beta_j\neq0)$
and $\beta_*$.
\end{longlist}\vspace*{-4pt}
\hrulefill
\caption{The steps in the HARNESS algorithm.}\label{figharness}
\end{figure}
The HARNESS bears some similarity to
POSI [\citet{Beretal13}]
which is another inference method for model selection.
They both eschew any assumption that the linear model
is correct.
But POSI attempts to make inferences that are valid over all
possible selected models while the HARNESS restricts
attention to the selected model.
Also, the HARNESS emphasizes predictive inferential statement.
Here is an example using the wine dataset.
(Thanks to the authors for providing the data.)
Using the first half of the data,
we applied forward stepwise \mbox{selection} and used
$C_p$ to select a model.
The selected variables are
Alcohol, Volatile\_Acidity, Sulphates,
Total\_Sulfur\_Dioxide and pH.
A 95 percent confidence interval
for the predictive risk of the null model is
(0.65,0.70).
For the selected model, the confidence
interval for $R$ is
(0.46,0.53).
The (Bonferroni-corrected) 95 percent confidence intervals for the $R_j$'s
are shown in the first plot of Figure~\ref{figwine}.
The (Bonferroni-corrected) 95 percent confidence intervals for the
parameters of the projected model
are shown in the second plot in Figure~\ref{figwine}.
\begin{figure}
\includegraphics{1175ef02.eps}
\caption{Left plot: confidence intervals for $R_j$.
Right plot: confidence intervals for projected parameters.
From top down the variables are
Alcohol, Volatile\_Acidity, Sulphates,
Total\_Sulfur\_Dioxide and pH.}
\label{figwine}
\end{figure}
\section{The value of data-splitting}
Some statisticians are uncomfortable with data-splitting.
There are two common objections.
The first is that the inferences are random: if we repeat the procedure
we will get
different answers.
The second is that it is wasteful.
The first objection can be dealt with by
doing many splits and combining the information appropriately.
This can be done but is somewhat involved and will be described elsewhere.
The second objection is, in my view, incorrect.
The value of data splitting is that leads to
simple, assumption-free inference.
There is nothing wasteful about this.
Both halves of the data are being put to use.
Admittedly, the splitting leads to a loss of power
compared to ordinary methods
\textit{if the model were correct}.
But this is a false comparison since we are trying
to get inferences
without assuming the model is correct.
It is a bit like saying that nonparametric function
estimators have slower rates of convergence than parametric estimators.
But that is only because the parametric estimators invoke stronger assumptions.
\section{Conformal prediction}
Since I am focusing my discussion on
regression methods that make weak assumptions,
I would also like to briefly mention
Vladimir Vovk's theory of
conformal inference.
This is a completely distribution-free, finite sample
method for predictive regression.
The method is described in
\citet{VovGamSha05}
and
\citet{VovNouGam09}.
Unfortunately,
most statisticians seem to be unaware of this work
which is a shame.
The statistical properties (such as minimax properties)
of conformal prediction
were investigated in
\citet{Lei}, \citet{LeiWas}.
A full explanation of the method is beyond the scope of this
discussion but I do want to give the general idea
and say why it is related the current paper.
Given data
$(X_1,Y_1),\ldots,(X_n,Y_n)$,
suppose we observe a new $X$
and want to predict $Y$.
Let $y\in\mathbb{R}$ be an arbitrary real number.
Think of $y$ as a tentative guess at $Y$.
Form the augmented data set
\[
(X_1,Y_1),\ldots, (X_n,Y_n),
(X,y).
\]
Now we fit a linear model
to the augmented data
and compute residuals
$e_i$ for each of the $n+1$ observations.
Now we test
$H_0\dvtx Y=y$.
Under $H_0$,
the residuals are invariant under permutations
and so
\[
p(y) = \frac{1}{n+1}\sum_{i=1}^{n+1} I\bigl(
|e_i| \geq|e_{n+1}|\bigr)
\]
is a distribution-free $p$-value for $H_0$.
Next, we invert the test:
let
$C = \{y\dvtx p(y) \geq\alpha\}$.
It is easy to show that
\[
\mathbb{P}(Y\in C) \geq1-\alpha.
\]
Thus, $C$ is distribution-free, finite-sample
prediction interval for $Y$.
Like the \mbox{HARNESS},
the validity of the method does not depend
on the linear model being correct.
The set $C$ has the desired coverage probability no matter what
the true model is.
Both the HARNESS and conformal prediction use the linear model
as a device for generating predictions but neither
requires the linear model to be true
for the inferences to be valid.
[In fact, in the conformal approach, any method of generating residuals
can be used.
It does not have to be a linear model. See \citet{LeiWas}.]
One can also look at how the prediction interval $C$
changes as different variables are removed.\vadjust{\goodbreak}
This gives another assumption-free method to
explore the effects of predictors in regression.
Minimizing the length of the interval
over the lasso path can also be used as a
distribution-free method for choosing the regularization
parameter of the lasso.
On a related note,
we might also be interested in
assumption-free methods for the related task
of inferring graphical models.
For a modest attempt at this,
see \citet{WasKolRin}.
\section{Causation}
LTTT do not discuss causation.
But in any discussion of the assumptions
underlying regression,
causation is lurking just below the surface.
Indeed, there is a tendency to conflate
causation and inference.
To be clear: prediction, inference and causation
are three separate ideas.
Even if the linear model is correct,
we have to be careful how we interpret the parameters.
Many articles and textbooks describe $\beta_j$
as the change in $Y$ if $X_j$ is changed, holding the other covariates fixed.
This is incorrect.
In fact, $\beta_j$ is the \textit{change in our prediction of} $Y$
if $X_j$ is changed.
This may seem like nit-picking but this is the
very difference between association and causation.\looseness=-1
Causation refers to the change in $Y$
as $X_j$ is changed.
Association (prediction) refers to the change
\textit{in our prediction of} $Y$
as $X_j$ is changed.
Put another way,
prediction is about $\mathbb{E}(Y| \mbox{ observe } X=x)$ while
causation is about $\mathbb{E}(Y| \mbox{ set } X=x)$.
If $X$ is randomly assigned, they are the same.
Otherwise, they are different.
To make the causal claim, we have to include
in the model, every possible confounding variable
that could affect both $Y$ and $X$.
This complete causal model has the form
\[
Y = g(X,Z) + \varepsilon,
\]
where $Z=(Z_1,\ldots,Z_k)$
represents
all confounding variables in the world.
The relationship between $Y$ and $X$
alone is described as
\[
Y = f(X) + \varepsilon'.
\]
The causal effect---the change in $Y$
as $X_j$ is changed---is given by $\partial g(x,z)/\partial x_j$.
The association (prediction)---the change in our prediction of $Y$
as $X_j$ is changed---is given by $\partial f(x)/\partial x_j$.
If there are any omitted confounding variables,
then these will be different.
Which brings me back to the paper.
Even if the linear model
is correct,
we still have to exercise great caution in interpreting
the coefficients.
Most users of our methods
are nonstatisticians and are likely to interpret
$\beta_j$ causally no matter how many warnings we give.
\section{Conclusion}
LTTT have produced
a fascinating paper
that significantly advances
our understanding of
high-dimensional regression.
I expect there will be a flurry
of new research inspired by this paper.
My discussion has focused on the role of
assumptions.
In low-dimensional models,
it is relatively easy to
create methods that make few
assumptions.
In high-dimensional models,
low assumption inference is much more challenging.
I hope I have convinced the authors that
the low assumption world is worth exploring.
In the meantime, I congratulate the authors
on an important and stimulating paper.
\section*{Acknowledgments}
Thanks to Rob Kass, Rob Tibshirani and Ryan Tibshirani
for helpful comments.
|
\section{Introduction}
\subsection{Cool star activity}
Magnetic activity drives high-energy processes in the atmospheres of stars. Powerful releases of energy (flares) are regularly observed and display plasma temperatures of several 10$^7$ K, reaching up to $\sim 10^8$ K in extreme cases (see e.g., G\"{u}del et al., 1999; Maggio et al., 2000; Favata et al., 2000; Franciosini et al., 2001). Such processes give rise to observable effects in the microwave region, and while the high-energy tail of thermal electrons has been observed in hard X-rays (Pallavicini \& Tagliaferri, 1999; Favata \& Schmitt, 1999; Franciosini et al., 2001), evidence on the high-energy non-thermal electron population has generally been scant.
One of the important facts to emerge from soft X-ray (SXR) observations during the Skylab era was the recognition that magnetic loops constitute the basic `building blocks' of coronal structure (Rosner et al., 1978). Using
the solar analogy, it became clear by 1980 that loops play a basic role in coronal energetics in both active and quiescent phases (e.g. Pallavicini et al., 1977, Kane et al., 1980).
Regarding flares, according to the evaporation model (cf.\ Antonucci et al., 1984), beams of fast electrons travel along loops and deposit their energy towards the lower chromospheric material. Heated plasma may subsequently expand into the loop structures. This hot plasma cools through conduction and radiation, but more than one emission or absorption mechanism is likely to be involved (Dulk et al., 1986, Bastian \& Gary, 1992, Alissandrakis et al., 1993).
In the cm range, however, the dominant emission usually involves the gyrosynchrotron process from loops.
The solar model is often used in trying to understand the mechanism and location of flares in active stars. A non-thermal electron population is thought to be present in the impulsive phase of flares (Doschek, 1972) and this is relevant to the power-law index of the electron energy distribution and its lower energy bound (Phillips \& Neupert, 1973), used in microwave data analysis. Hard X-rays ($>$10~keV) should be emitted during the initial stages of the flare. Softer X-rays become detected as high energy electrons impinge on denser (`chromospheric') layers as the flare develops, giving rise to thermal bremsstrahlung. Complementary microwave observations should evidence the non-thermal electron source, giving information on its energy spectrum and possible changes during the course of a flare. In fact, radio data have sometimes pointed to a very strong non-thermal electron component (e.g., Lim et al., 1994).
The present study concentrates much of its in-depth analysis
on such microwave data from the active star AB Doradus in an effort to gain clearer insights into
how such large energy releases occur.
Within the two days of combined X-ray and radio data we discuss, the latter showed higher signal to noise (S/N) ratio for typical integration intervals of several minutes, particularly in the higher frequency bands.
The multi-wavelength campaign that stimulated the present work was originally aimed at examining, in particular, the initial population of highly accelerated (non-thermal) electrons. While very hard, non-thermal X-rays would be too weak to detect in quiescence, they were calculated to be picked up by Suzaku's sensitive detectors during the onset of large flares. Simultaneous radio and X-ray observations then have the possibility to reveal the `Neupert effect' in flares. This concerns the correlation between radio and soft X-ray or UV flare emissions (Neupert, 1968), in which the short wavelength light curve of radiatively cooling plasma tends to follow the integral of the radio emission. This can be interpreted in terms of `loss cone' gyrosynchrotron-emitting electrons encountering denser chromospheric layers, whereupon X-ray emission follows the radio decline with a short time lag. The Neupert effect has been observed in several stars using radio and optical U-band data (Hawley et al., 1996; G\"{u}del et al., 1996; 2002a,b), and also radio and X-ray data (Osten et al., 2004).
This multiwavelength surveillance has allowed checking of rotational modulation effects. In particular, AB Dor is associated with long-lived centres of surface activity, noticed by persistent broadband (optical) maculation. The relationship of electron energization to such activity centres can be explored. In this way, optical photometry and high-resolution spectroscopy, observed over intervals around the time of the X-ray and radio observations, supplement the overall picture of this intriguing star.
\subsection{AB Doradus}
AB~Dor (HD~36705, K1V-type, $V$ = 6.93 mag, $M$ = 0.76 $M_{\odot}$) is a well-studied magnetically active young star that has been observed over a wide range of wavelengths. The main target is the primary component of a small multiple system containing at least 4 stars. Historically, AB Dor A was regarded as the primary of the visual binary Rst 137 (separation around 7.5 arcsec). More recently both AB Dor A and B were shown to be both binaries, involving low-mass companions (see e.g.\ Janson et al., 2007, for a review giving modern parametrization).
The K-type main star, at 14.9~pc distance (Perryman et al., 1997), displays numerous strong flares on time scales from minutes to weeks, with two extreme cases having integrated X-ray fluxes of $\sim$4$\times$10$^{-9}$erg cm$^{-2}$ s$^{-1}$, i.e.\ about 60 times the quiescent level. Observed flaring occurs at a mean rate of about one per day. Peak temperatures of around 110$\times$10$^6$~K (Figure~1 in Maggio et al., 2000) were inferred. This main star, AB~Dor Aa, has the relatively fast mean rotation of 1.40$\times$10$^{-4}$ radians~sec$^{-1}$ ($\sim$12.3~h period) and was long associated with surface `spots' that allow not only the mean rotation, but also differential rotation and cyclic behaviour to be examined
(J\"{a}rvinen et al., 2005).
The conclusion of Rucinski (1983) and Innis et al.\ (1986) that AB Dor Aa (subsequently, usually just `AB Dor') may be a pre-main sequence star was questioned by Micela et al.\ (1997). Still, AB Dor turns out to be very probably near to the zero-age main sequence (ZAMS). Zuckerman et al.'s (2004) checking of the galactic space motions of the Moving Group that contains AB Dor indicates an age of about 50 My. The youth of the system was confirmed in part by lithium abundance studies (Rucinski, 1982, Hussain et al., 1997).
There are various markers of AB Dor's magnetic activity. As well as the short-term `starspot' behaviour, photometric studies over long periods have shown the average brightness goes through cycles over the years. Observations over the period 1978-88, showed the star's average brightness to decrease steadily, reaching a minimum by the end of this interval. Subsequently, increases over the years 2000-2004 were observed and a maximum level reached (Amado et al., 2001; J\"{a}rvinen et al., 2005). This longer-term variation has been seen as an indicator of solar-type cycles.
Another indicator of AB Dor's magnetic fields is its active chromosphere. This is marked by the presence of strong CaII and H alpha emission (Bidelman \& MacConnell, 1973; Vilhu et al., 1987; Budding et al., 2009). As well, the active corona gives rise to strong and variable radio and X-ray radiation (Pakull, 1981; Slee et al., 1986), although
the evidence from different spectral regions is not
clear-cut regarding indications of activity (K\"{u}rster et al., 1997).
The presence of enlarged coronal structures on AB Dor has been discussed at least since the
giant loop model of Mestel and Spruit (1987), as well as the observational confirmation of emitting material linked to plasma at heights of order
several stellar radii above the photosphere by Collier Cameron and Robinson (1989). Loop sizes
were inferred from analysis of the fading of X-ray flares observed with the BEPPOSAX satellite by Maggio et al.\ (2000), but flaring-loop heights were, in that case, taken to be somewhat smaller ($H\lesssim$ $0.3 R_{\rm star}$), since the flares were observed to continue through all the rotation cycle without showing any eclipse effects, so that their source should not be far above the polar regions. The rotation axis is believed to be inclined at $\sim$60$^{\circ}$ to the line of sight. Other methods have tended to
support this scale of size for X-ray emitting structures
(Sanz-Forcada et al., 2003; Hussain et al., 2005.)
A comprehensive assessment of coronal energization, involving X-ray observations repeatedly collected over a three year period from the XMM-Newton and Chandra satellites was given by Sanz-Forcada et al.\ (2003)
and summarized by Sanz-Forcada et al.\ (2004).
Loop sizes were inferred to come in a distribution of sizes, but mostly relatively small compared to the stellar radius. The high energy tail of electron energy distributions
was found to vary significantly with the general activity state of the star. Hussain et al.\ (2007) combined evidence from the rotational modulation of the X-ray flux
with high dispersion ground spectroscopy to
produce detailed maps of likely emission configurations.
Possible geometries of coronal emission regions were also considered by Lim et al.\ (1994), who
examined radio data-sets comparable to ours.
Evidence on flare region geometry is, as a whole, rather incomplete and often somewhat indirect, however.
Modelling tends to involve many adjustable parameters,
implying over-determined curve-fittings and ambiguous results.
AB Dor has been the subject of a number of previous
satellite-based studies: for example, Rucinski et al.\ (1995); Mewe et al.\ (1996); K\"{u}rster et al.\ (1997);
Vilhu et al.\ (1998); Schmitt et al.\ (1998);
Maggio et al.\ (2000) and others, as discussed by
Sanz-Forcada et al.\ (2003).
Repetitive phase-linked phenomena have demonstrated
the presence of localized inhomogeneities,
with effects involving large-scale photospheric magnetic field concentrations,
through to chromospheric and transition region condensations and on to hot coronal plasma volumes
typically in the range of several to a few tens of millions of K -- occasionally more extreme.
Recently, Lalitha \& Schmitt (2013) have considered the
relationship between trends of X-ray emission and
optical mean flux, associated, tentatively, with a 17 y period.
Lalitha \& Schmitt found that the high basal level of X-ray emission associated with the strong activity of AB Dor A
does vary in response to the expected long-term cycling,
but the amplitude of this background emission is not so great
compared with other known cases, in particular the Sun.
So although the solar example continues to underlie many of the concepts, as with quiescent and active conditions,
the situation of AB Dor can involve structures and energies at least a
couple of orders of magnitude greater. Alternative scenarios can be applied to the emission mechanisms, and further clarification of this subject forms a strong motivation for the present work.
The cyclic, quasi-repetitive character of AB Dor's radio behaviour, over long periods of time, confirms the existence of structures maintaining the same longitude. The sources are explained in terms of field concentrations that slowly migrate over a number of years. Lim et al.\ (1994), from studying radio emission peaks and simultaneous optical data, deduced the emission to arise from close to the spotted areas. Many repeated phenomena from photosphere, chromosphere and corona thus point to the presence of `active longitudes'.
Notwithstanding the foregoing argument on flaring loop heights, there is also evidence of very extended coronal condensations. For example, the FUSE (Far Ultraviolet Spectroscopic Explorer) and HST (Hubble Space Telescope) have observed transition region lines (CIV 1548\AA, SiIV 1393\AA\ and OVI 1032\AA, $T \sim10^5$ K), with speeds of up to 270 km s$^{-1}$. An extended region of optically thin plasma, with high-level co-rotation, at heights up to 2.6 $R_{\rm star}$ was then suggested (Brandt et al., 2001; Ake et al., 2000).
Doppler Imaging techniques were often applied to AB Dor (K\"{u}rster et al., 1994; Unruh and Collier Cameron, 1994; Donati and Collier Cameron, 1997; Donati et al., 1999, Hussain et al., 2002). These studies of the surface distribution of active regions have demonstrated the persistence of two dominant spot longitudes, but at different latitudes. Donati and Collier Cameron's (1997) account of AB Dor from optical cross-correlation techniques concluded that the scale of differential rotation was approximately 40 times greater than that of the Sun. This would then suggest a preferred latitude for spots (in a given hemisphere), or the differential rotation would soon produce different phases for the recurrent maculation wave on which the ephemeris and consequent longitude system is based.
\section{Observations}
\subsection{X-ray observations}
Launched in 2005, Suzaku is engaged in observations of astronomical X-ray sources that form part of a joint programme of the Japanese Institute of Space and Astronautical Science (ISAS-JAXA) in collaboration with NASA-GSFC and other organizations (cf.\ Mitsuda et al., 2007). Three of Suzaku's four X-Ray detectors of the imaging spectrometer (XIS) were operable during our allocations. A hard X-ray detector (HXD) sensitive over the energy range 0.2-600 keV with low background and an angular resolution that should allow application to research on brighter flare stars
was also available.
A Suzaku proposal for observations of AB Dor was thus made in early 2006 with an aim to research
the high-energy non-thermal electron population, particularly during the initial flare impulse. Possible detection of very hard, non-thermal X-rays could allow better understanding of the triggering mechanism.
The X-ray observations, made simultaneously with the radio ones, were carried out in November 2006 and January 2007.
Unfortunately, the signal from the HXD during these observations did not permit reliable analysis (Arnaud, private communication). However, the XIS (intermediate energy) data show interesting effects.
The XIS has a field of approximately 18 arcmin square covered by the three available detectors XIS0, 1, and 3.
The angular resolutions of the Suzaku telescopes range from 1.8 to 2.3 arcmin half-power diameter. This resolution does not significantly depend on the energy of the incident X-rays in the available range, but it entails that all four known components of the
multiple star are covered by the acceptance beam.
The archived Chandra data from 2004 (http://cda.harvard.edu/ pop/), has much higher angular resolution
and it shows two separate sources with both a 52 ks exposure using the HETG, and an 85 ks exposure with the LETG grating, in the positions of AB Dor A and B on the sky.
Aperture photometry of these images shows the AB Dor A source
(as enclosed within a 1.7 arcsec radius aperture)
to have an average flux density of 83.3\% of the sum of the two sources with the HETG instrument, and 87.8\% with the LETG.
Since these exposures cover time intervals of the order of a day, there is a good chance that they
would have included a flare, but mostly the output would be dominated by accumulated quiescent
emission from the two heated coronae in the proportions 5.7 to 1, for AB Dor A to B.
The simultaneous, higher angular resolution, radio observations show that three large flares occurred from AB Dor A during the
joint coverage. The X-ray cameras registered distinct
flares around the same times. Further insights are gained when the X-ray and microwave observations are superposed, as discussed in Section 4.
\begin{figure*}
\begin{center}$
\begin{array}{c}
\hspace{0cm}
\includegraphics[scale=1.1, angle=0]{fig1an2.png} \\
\hspace{0cm}
\includegraphics[scale=1.1, angle=0]{fig1bn2.png}
\end{array}$
\caption{Light curves of AB Dor collected by the Suzaku XIS detectors; top: for Nov 21-22, 2006; and bottom for Jan 8, 2006. These X-ray raw data from the intermediate range cameras integrate photon energies between 0.3 to 10 keV. The $y$-axis count units are linearly proportional to the received flux.}
\label{AB_Dor_f1}
\end{center}
\end{figure*}
The X-ray data were taken from the HEASARC archive (NASA, cf.\ White, 1992). HEASoft reductional software, including {\sc FTOOLS} and {\sc XANADU}, is similarly available. In Figure~1 we show the integrated XIS raw data over the range 0.3 to 10 keV from the two observation dates.
The original data were collected on 2006 Nov 21-22 (HJD 2454060-1; 2h 43m 51.9s until 11h 01m 35.2s, i.e.\ $\sim$116000 s) and on 2007 Jan 8-9 (HJD 2454108-9; 2h 46m 25.2s until 10h 48m 3.7s, i.e.\ $\sim$115000 s).
\begin{figure*}
\vspace{0cm}
\begin{center}
\hspace{-1cm}
\includegraphics[scale=0.27, angle=90]{fig2.png}
\caption{{(top) AB Dor, 21/22 November 22, 2006, Suzaku X-ray spectra derived from integrations with the XIS1 (upper trend of points), XIS0 and XIS3 (middle and lower) cameras, respectively. The left panel corresponds to the quiescent flux,
the right to flaring. (bottom) The same arrangement but for 8/9 January, 2007. The Fe XXV emission at around 6.7 keV feature is only apparent at times during flares.}}
\label{AB_Dor_njspec}
\end{center}
\end{figure*}
Individual 30 s integrations with the separate cameras during quiescent intervals show a typical scatter of about 12\%\ of measured levels.
This implies that time averaged values covering a few hours of quiescence would reduce the error of a representative mean to a few percent. We can
check this in the foregoing averaged values, i.e.\
$f_{\rm av} = 4.71 \pm 0.04 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$, for the Nov data and
$f_{\rm av} = 6.52 \pm 0.13 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$, for the more active Jan data.
The X-ray spectra matching these net fluxes, with their
modellings, are presented in Figure~2 (left).
\begin{figure*}
\begin{center} $
\begin{array} {c}
\hspace{-0.8cm}
\includegraphics[scale=1.2, angle=0]{fig3an2.png} \\
\hspace{-0.8cm}
\includegraphics[scale=1.2, angle=0]{fig3bn2.png}
\end{array}$
\caption{(top)AB Dor, 21/22 November, 2006, Suzaku X-ray observations for different energy ranges
(see Section 2.1 text). Data from the XIS1 camera (lowest energy range) are shown as the upper trend of points (red), with middle (green) and lower (blue) trends corresponding to the (higher energy) XIS3 and XIS0 cameras respectively (bottom). Related TBabs and modelling parameters are referred to in the text. The same arrangement but for 8/9 January, 2007.}
\label{AB_Dor_xfig5}
\end{center}
\end{figure*}
\begin{figure*}
\begin{center} $
\begin{array} {c}
\hspace{-0.8cm}
\includegraphics[scale=0.45, angle=-90]{novdelx1.png} \\
\hspace{-0.8cm}
\includegraphics[scale=0.45, angle=-90]{jandelx1.png}
\end{array}$
\caption{AB Dor, 21/22 November, 2006, spectral index variation corresponding to Figure~3 (top) and for 8/9 Jan, 2007 (bottom).}
\label{AB_Dor_xfig5a}
\end{center}
\end{figure*}
The XIS spectra were processed using {\sc XANADU}, particularly its {\sc XSPEC} spectral-fit program.
This calls on the Tuebingen-Boulder ISM absorption model {\sc TBabs}, which accounts for the intervening effects of the interstellar medium. Since the modelling is overdetermined by the available parameter-set, our approach was minimalistic
in trying to find the smallest number of independent parameters
consistent with the data. The processing involves use of the {\sc MEKAL} (MEwe-KAastra-Liedahl) code, constructing multi-temperature plasma emission models (Kaastra et al., 1996). This analysis was carried out by one of us (NE) whilst on study leave at the Armagh Observatory.
X-ray fluxes from stellar coronae are characterized by a distribution of emission measures (EMD)
for the various contributing species to the net emission.
The flux normally shows a distinct maximum, with a
corresponding brightness temperature $T_{\rm max}$,
typically in the order of 10 MK.
The brightness temperature $T$ can be related to
the EMD through a proportionality of the form ($T$/$T_{\rm max})^{\alpha}$. {\sc MEKAL} model parameters include $\alpha$ and $T_{\rm max}$, $nH$, abundances of the elements, and a normalization parameter. In the present cases we derived (from optimal curve-fitting) $T_{\rm max} = 2.588 \pm 0.039$ (keV) and $T_{\rm max} = 3.129 \pm 0.046$ (keV, with s.d.\ error measures here and elsewhere), i.e.\ about 9.3 and 11.2$\times$10$^6$ K, for the November and January allocations respectively. Acceptable fits were
found with $\alpha$ set to 0.1, the corresponding low variation of the EMD with brightness temperature,
indicative of dense and compact conditions
for the spectral region observed.
Further details on the reduction and modelling procedures were given by Erkan (2012).
The spectral analyses covered all three cameras' data and the results were cross-checked for self-consistency. For the range 0.3 to 10 keV energy,
{\sc TBabs} modelling of the quiescent levels gave the following flux
averages:
\vspace{2ex}
\begin{tabular}{c}
21/22 November, 2006 \\
$f_{\rm XIS0} = 4.69 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$;\\
$f_{\rm XIS1} = 4.77 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$;\\
$f_{\rm XIS3} = 4.65 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$;
\end{tabular}
\begin{tabular}{c}
8/9 January, 2007 \\
$f_{\rm XIS0} = 6.34 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$;\\
$f_{\rm XIS1} = 6.55 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$;\\
$f_{\rm XIS3} = 6.70 \times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$;
\end{tabular}
\vspace{2ex}
The high energy range tail of the X-ray spectrum
is characterized
by an index $\delta$, where $F_{\nu} \propto \nu^{\delta}$ (Dulk, 1985) and $\delta \approx -3$ is often observed in practice\footnote{Dulk (1985) actually wrote $F_{\nu} \propto \nu^{-\delta}$.}. Our results are similar to those found for other active cool stars
(e.g.\ Osten et al., 2002), showing a peak at about 1 keV and a high energy tail with a power-law slope close to --3 in the range 1-5 keV. Energy resolution beyond 5keV is less distinct, although the 6.7 keV Fe emission feature becomes clear during flares: at other times it was not detected, as shown in Figure~2.
Data analysis involves the so-called Good Time Intervals (GTI), selected
according to investigator preference
but generally in accordance with suitable signal levels. These intervals may be extended over times of lesser scatter,
or perhaps flaring episodes.
Figure~3 presents reduced observations of AB Dor. GTIs are here ascribed simply
in accordance with the more continuous on-time integrations of the source on the two observation dates.
Flux-weighted averages for the three bandpasses (low, medium
and high, say)
in Figure~3 correspond to 2.42, 6.91 and 14.7$\times$10$^{16}$ Hz as their effective frequencies. The low and medium energy ranges allowed clear measures of the X-ray incident flux throughout the monitoring. Spectral indices can then be derived and
compared with the average spectra shown in Figure~2.
Data from GTI regions 8-10 in Nov 06 and 9-17 in Jan 07 correspond to times when prominent flares were observed.
Flux levels in the highest frequency channel (5-10 keV) had
very low signal strength throughout the monitoring, with
S/N ratios for individual integrations of typically $\sim$1.5, during quiescence, rising to about 3 near the flare peaks.
These low levels of the high energy signal are consistent,
when integrated over the selected time intervals,
with the power-law form of the high energy tails in Figure~2.
In the flare regions, the highest energy channel shows distinctly more noticeable flux in Figure 3. These flux increases across the medium-high range
are indicative of spectral hardening. This can be seen in Fig 3 with GTIs 8 and 9 on Nov 06, but more so near the flare peaks in GTIs 9 and 12 of Jan 07, particularly the latter.
Since the flux distributions on either side of the maximum
tend to a power-law-like form (albeit with
different indices)
flux-versus-frequency integrals tend to be dominated
by one of the limiting ordinate values.
This means that the low energy range fluxes
can be scaled by a suitable constant to support estimates of the spectral power law index for the high energy tail.
This was checked
empirically, and the scaling factor 3.2 was found to
allow the mean value of $\delta$ from the low to medium flux ratio to agree with that for the medium to high range. In this way, the standard deviation of the $\delta$ values
were generally reduced from about 0.25 to 0.08. The corresponding
$\delta$ values, with their s.d.\ errors, are shown
in Figure~4. The sought differential variation of $\delta$ thus reflects the clear relative intensification of the
medium flux measures during flares.
$\delta$ can be seen to approach --2.7 at the peaks in Fig~4. Noticeable also about the second flare of the Jan data is that this hardening occurs at the initial X-ray peak, but during the subsequent microwave peak and second slower enhancement the X-ray emission is softer.
This corresponds to the classic Neupert effect, and the pattern is consistent with an initial high-energy triggering followed by a gyrosynchrotron flare with a later thermal cooling associated with electron precipitation.
This point has interesting consequences for our appraisal of flare behaviour, and we return to it in the later discussion.
\subsection{Radio observations}
\subsubsection{ATCA Data}
Preliminary forms of AB Dor's radio light curves were given by Budding et al.\ (2009), mainly with the aim of showing that the various activity manifestations at different wavelengths were probably linked to identifiable centres of activity on the primary star.
However, that was without detailed backgrounding of the observational circumstances and data analysis that we now present in this and following
sections of the present paper. In fact, the idea of repetitive phase-related enhancements in AB Dor's radio emission is not new. It was emphasized in the the study of Lim et al.\ (1994), where radio light curves showing a two-peaked structure that aligned with optically identified photospheric centres of activity was evidenced. Lim et al.'s observations appear to have found the star in a more active condition than the present authors, however. The peak intensities were significantly higher (by factors of typically $\sim$5) than those we discuss in what follows. The emission turnover frequencies were also generally lower: our 8.6~GHz peaks being noticeably lower than those at 4.8~GHz. At the time of these observations the Australia Telescope Compact Array (ATCA) was an East-West synthesis telescope of total length 6~km operating through cm to mm wavelengths.\footnote{The facility was significantly upgraded in 2010 with the addition of the Compact Array Broadband Backend
(CABB), enabling a 2 GHz observing bandwidth.} In the cm bands the array was then capable of simultaneous dual-frequency observations. In this experiment we operated at 1.384/2.368~GHz and 4.80/8.64~GHz in an attempt to explore the physical conditions responsible for the star's gyrosynchrotron emission over a large range of heights in its corona.
\begin{figure*}
\vspace{0cm}
\begin{center} $
\begin{array} {cc}
\hspace{-1cm}
\includegraphics[scale=0.40,angle=-90]{21x.png} &
\includegraphics[scale=0.40,angle=-90]{8x.png} \\
\hspace{-1cm}
\includegraphics[scale=0.40,angle=-90]{21c.png} &
\includegraphics[scale=0.40,angle=-90]{8c.png}
\end{array}$
\caption{
Robustly weighted contour maps of AB Dor A and B (northern component), using the uv data from 15 baselines. The maps on the left are for 21 Nov 2006; on the right
8 Jan 2007. The 8.64 GHz results are at the top; 4.80 GHz below.
Map (a) has highest and lowest contour levels of 2.62 and 0.20 mJy/beam, and the system noise
over a large clear area around the source is 51 $\mu$Jy/beam. The beam axes are 3.36 and 2.84 arcsec at PA 71.2 deg.
Map (b) has highest and lowest contour levels of 3.37 and 0.15 mJy/beam, and corresponding noise 47 $\mu$Jy/beam. The beam axes are 4.45 and 3.51 arcsec; PA 72.0 deg.
Map (c) has corresponding contour levels of 2.85 and 0.25 mJy/beam and noise
65 $\mu$Jy/beam. The beam axes are 4.82 and 4.52 arcsec; PA --23.6 deg.
Map (d) has corresponding contour levels of 4.08 and 0.33 mJy/beam and noise
62 $\mu$Jy/beam. The beam axes are 8.56 and 8.01 arcsec; PA - 23.6 deg.}
\label{AB_Dor_contour_map}
\end{center}
\end{figure*}
Ideally, we require an array configuration providing angular resolution much less than the $\sim$8~arc~sec separation of AB Dor A from its M3.5V visual companion at the four frequencies of interest, in order to identify the individual contributions of the two stars.
In the event, we were granted a 1.5~km configuration (5 antennae) on the Nov 21/22 allocation, and a 750~m configuration on the Jan 8/9 allocation. Both these configurations (especially the 750~m one) were less than optimal for the task, but we were able to obtain satisfactory resolution of the stars' emissions at 4.800 and 8.640~GHz. Reductions have been performed using the MIRIAD software package (Sault et al., 1995), which has become well-adapted for application to ATCA data.
The sampling interval for our observations was 10s and samples were accumulated for 14.2 h in Nov, 2006 and 11.4 h in Jan, 2007. The two pairs of bands were alternately recorded in 25 m integrations and calibrated using MIRIAD. After we mapped the two sets and discovered evidence for correlated activity, we further divided each 25 m integration in the C/X bands into 5 m integrations in a search for finer temporal structure, and to assist in deriving accurate correlation functions.
Analysis of the 1.384 and 2.368~GHz data was further complicated by the presence of persistent, strong interference, particularly on the ten shorter baselines. We found it, eventually, impracticable to produce useful light curves at the longer wavelengths, although resolved images were obtained at the longer wavelengths from integrations over the total allocation intervals.
Lack of angular resolution with the more compact array of Jan 8, 2007 meant that the usual mapping tasks in MIRIAD gave problems for the separation of AB Dor A and B on all short time integrations. Maps weighted with the `robust' option show an almost complete resolution at X band, however, and a usable result at C band.
It was decided to use this option to produce all four maps shown in Fig~5. The resolution is then good enough to permit UVFIT to converge to consistent flux and rms values. The outer contour for the Jan 8 data (especially at X-band) shows the wavy structure compatible with a significant gap in uv coverage between the 5 compact baselines and antenna 6.
To attain better resolution at L/S on both observing days we tried the 5 longest baselines to antenna 6, resulting in a marked reduction in the S/N of the 25 minute integrations. Our L~band data were then found still contaminated by terrestrial interference and we did not analyse them further. We also do not present the S~band light curve in this paper, since it is not in the same S/N category as the shorter wavelength data.
We used the MIRIAD task UVFIT on both 25-m and 5-m integrations for the C and X bands. UVFIT fits a model to the uv data, using accurate offsets from the phase centre for the A and B components. These offsets were derived from cleaned contour maps of the 5-baseline data to antenna 6, using the task IMFIT. Our final weighted means of the four independent sets of offsets at each frequency are $\Delta x = 0.430$ arcsec and -–1.797 arcsec; and $\Delta y = +9.865$ and +18.625 arcsec for components A and B, respectively. The errors in these measurements are $\pm$0.035 and $\pm$0.065 arcsec, respectively.
These offsets were used in deriving the flux densities.
We have made naturally-weighted contour maps at 4.80 and 8.64~GHz from all the 25~min.\ integrations available on each of the three observation periods (two in November 2006 and one in January 2007). Maps were made at each of the following combinations of baselines; (1) the total of 15 baselines; (2) the 10 compact baselines; and (3) the 5 long baselines. A comparison of the maps allowed us to select the combination that would permit measurement of the flux densities from the individual 25~min.\ integrations with the best combination of sensitivity and angular resolution. The best images allowed us to check the used angular offsets of the stellar images from the field centre (see Figure~5).
The radio position corresponding to the best integrated image yields AB Dor A at 2006.894 to have RA: 05 h 28 m 44.924 s +/- 0.007; and dec:
--65 deg 26 arcmin 53.910 arcsec +/- 0.039. This may be compared with the ICRS position for 2000.0, i.e.
RA: 05 h 28 m 44.8280 s +/- 0.001; and dec: --65 deg 26 arcmin 54.853 arcsec +/- 0.004. Mean proper motions that could be derived from this pair of positions and 6.894 y interval are different from the HIPPARCOS-based (van Leeuwen, 2007) values listed by SIMBAD, however. This may well be related to orbital motion within this multiple system, as noted in a similar context by Martin and Brandner (1995), from whose data over the 65 year interval 1929-94 proper motions of +51$\pm 5$ and +130$\pm 5$ mas y$^{-1}$ in RA and dec respectively can be found. These are closer to the +87$\pm 3$ and and +137$\pm 3$ results of our measures, though still well outside the measurement errors from the SIMBAD figures of 33.2$\pm 0.4$ and 150.8$\pm 0.7$. But, from the orbit of the AB Dor A-C system shown by Nielsen et al.\ (2005), it is clear that
orbit-related proper motion changes of the order of tens of mas y$^{-1}$ can be reasonably expected from year to year.
The short-interval HIPPARCOS values may thus be significantly affected in this way.
Maps of the source are shown in Figure~\ref{AB_Dor_contour_map}. A log of the observations and the resulting mean flux densities for each night are given in Table~\ref{table:statistics}, which also contains the mean spectral indices
$\alpha$, with the spectral index defined by S($\nu$) = C$\nu^{\alpha}$. The polarized flux in this table
refers to the circular (V) component.
\begin{table*}
{\scriptsize
\centering
\begin{minipage}{100mm}
\caption{ATCA observation log and summary of data.}
\label{table:statistics}
\begin{tabular}{ @{}ccccccccc@{} }
\hline
Date & Dur.\ & Star & Freq.\ &Mean &Pol'ized& \multicolumn{3}{c}{\mbox Spectral index }\\
Config. & hr& & GHz & Total & Frac.\ & $\alpha^{1}$: & $\alpha^{2}$: & $\alpha^{1}-\alpha^{2}$\\
& & & & Flux & V/I & 2.37--4.80 & 4.80--8.64 & \\
& & & & (mJy) & & GHz & GHz & \\
\hline
& & & & & & & & \\
21 Nov. 2006& 17.1 & AB Dor A &2.37 &3.20 (0.11)& 0.04 (0.10) &+0.27 (0.05) &--0.35 (0.03) & 0.62 (0.06) \\
1.5B & & & 4.80 & 3.88 (0.04) & 0.05 (0.06) & & & \\
& & & 8.64 & 3.15 (0.05) & 0.15 (0.07) & & & \\
&&&&&&&&\\
& & AB Dor B & 2.37 & 1.60 (0.11) & 0.33 (0.10) &+0.17 (0.10) & --0.22 (0.06) & 0.39 (0.12) \\
& & & 4.80 & 1.80 (0.04) & 0.30 (0.06) & & & \\
& & & 8.64& 1.58 (0.05)& 0.34 (0.07) & & & \\
\hline
& & & & & & & & \\
08 Jan. 2007 & 13.3 & AB Dor A & 2.37 & 3.65 (0.11) & 0.01 (0.16) &+0.12 (0.05) &--0.20 (0.03) &0.32 (0.06) \\
750A& & & 4.80 & 3.96 (0.05) & 0.07 (0.07) & & & \\
& & & 8.64 & 3.51 (0.05) & 0.12 (0.07) & & & \\
&&&&&&&&\\
& & AB Dor B & 2.37& 1.73 (0.11) & 0.17 (0.16) &--0.21 (0.10) &+0.18 (0.08) & --0.39 (0.13) \\
& & & 4.80 & 1.49 (0.05) & 0.25 (0.07)& & & \\
& & & 8.64 & 1.66 (0.05) & 0.25 (0.07) & & & \\
\hline
\end{tabular}
\end{minipage}
}
\end{table*}
\subsubsection{Flux Density Plots}
\begin{figure*}
\vspace{0cm}
\begin{center}$
\begin{array}{cc}
\hspace{-0.25cm}
\includegraphics[scale = 0.29, angle = -90]{lncn.png} &
\includegraphics[scale = 0.29, angle = -90]{lnxn.png} \\
\hspace{-0.25cm}
\includegraphics[scale = 0.29, angle = -90]{ljcn.png} &
\includegraphics[scale = 0.29, angle = -90]{ljxn.png}
\end{array}$
\end{center}
\vspace{-0.4cm}
\caption{Plots of flux densities corresponding to the midpoints of 25~min.\ integrations on 2006 Nov 21 (upper panels) and 2007 Jan 08 (lower panels) for AB Dor A and B. These points utilize data from all 15 baselines
at 4.80 and 8.64 GHz. Photometric phases, corresponding to the times shown on the lower axis, are displayed on the top axis, using Innis et al.'s (1988) ephemeris. The rms error measures for both components as derived by UVFIT were the same; they are omitted from the lower light curves to avoid possible confusion when the signals are close.}
\label{AB_Dor_flux_densities}
\end{figure*}
The flux densities derived with UVFIT for the AB Dor system are plotted in the montage of Figure~6 for the observing runs on 2006 Nov 21 and 2007 January 08. The date, frequencies and stars are identified in the panels. Each point in these plots represents the flux density at the mid-UT of a 25 min integration, and the error bars represents the rms residual of the fit. Clear, long-lasting increases for AB Dor A are seen on both dates, soon after the start of observations on Nov 21 and a few hours in and again towards the end of the run on Jan 8.
The 2.27 GHz data-set is suggestive of correspondence with these effects,
but it is much less clear against the relatively high scale of noise
for the reasons given before.
Table~\ref{table:flares} presents estimates of peak flare intensities (i.e. excess emission above the background) and their total durations at both higher frequencies. The angular resolution was sufficient to show that the M-type dwarf binary, AB Dor B, remained relatively quiescent on all three nights and it is clear that the fluxes measured for AB Dor A are not significantly influenced by the companion. The scale of noise in the measurements is provided by both UVFLUX and UVFIT. The relatively strong cross-correlation of the C and X-band data, discussed later, provides further confirmation of the reality of the signal excursions during flaring episodes. Base levels for these events can be reasonably set at 2.0 mJy for all the curves, except for the 4.80 GHz curve on Nov 21, when the base level appears closer to 2.5 mJy. The excess flare fluxes are based on these levels.
\begin{table*}
\vspace{-0.3cm}
{\scriptsize
\centering
\begin{minipage}{100mm}
\caption{Microwave flare statistics.}
\label{table:flares}
\begin{tabular}{ @{}ccccccccc@{} }
\hline
Date & Flare & Flare & \multicolumn{3}{c}{\mbox Peak Flux Density (mJy)}& \multicolumn{3}{c}{\mbox Spectral Index}\\
& & UT (h) & 2.37~GHz & 4.80~GHz & 8.64~GHz & $\alpha^{1}$: & $\alpha^{2}$: & $\alpha^{1}-\alpha^{2}$\\
& No & Duration & & & & 2.37--4.80 & 4.80--8.64 & \\
& & & & & & GHz & GHz & \\
\hline
& & & & & & & & \\
06 Nov. 21 & 1& 07:25--14:50&2.33 (0.49)& 3.24 (0.20)& 2.71 (0.21)& 0.47 (0.31)& --0.30 (0.17)& 0.77 (0.35)\\
& 2& 12:50--18:50& 1.06 (0.38)& 1.68 (0.16)& 1.37 (0.16)& 0.65 (0.53)& --0.35 (0.26)& 1.00 (0.59)\\
& 3& 20:00--$\ge$23:00& $\le$0.5 & 1.11 (0.16) & 0.69 (0.16) & $\ge$1.46 & --0.81 (0.47)& $\ge$ 0.67 (0.70)\\
&&&&&&&&\\
07 Jan. 08 & 1&10:00--15:50 &1.72 (0.34) & 4.23 (0.21) & 3.07 (0.22) & 1.28 (0.29) & --0.55 (0.15) & 1.83 (0.33) \\
& 2 & 16:50--$\ge$21:50& 2.48 (0.40) & 5.95 (0.26) & 5.31 (0.26) & 1.24 (0.24) & --0.19 (0.11) & 1.43 (0.26) \\
\hline
\end{tabular}
\end{minipage}
}
\end{table*}
\subsection{Optical Observations}
The broadband $B$ and $V$ photometry and high resolution, wide-range \'{e}chelle spectroscopy components of this campaign were discussed in more detail in Innis et al.\ (2008) and Budding et al.\ (2009). The latter paper presented an empirical model of surface activity, generally contemporaneous with the radio and X-ray data, consisting of four main centres. That paper
also gave reasons why surface features of the scale in question could be considered stable over timescales of a few months, in general.
One of the optical features discerned (a) appears to consist of a strong umbral region near, but somewhat above, phase zero ($\phi = 0.03$), that may also be related to a chromospheric enhancement at phase 0.95. This location agreement
strengthens the argument of applicability of not exactly time-coincident data to models across the complete spectrum (see also Lim et al., 1994).
The other three activity centres (b, c and d) correspond essentially to chromospheric enhancements, defined particularly in high dispersion Ca II emission K lines, are centred around phases (0.26, 0.58 and 0.84) respectively. We shall refer to this representation in interpreting the X-ray and radio coronal emission.
A more detailed combination of almost simultaneous optical and X-ray observations of AB Dor using the SemelPol polarimeter and UCLES spectrograph at the AAT and the Chandra satellite over a five day interval in late 2002
was carried out by Hussain et al (2007). It is of interest that there were qualitatively similar results regarding the small number of longitudinally confined major centres of activity at high latitudes. These connect optically identified photospheric and chromospheric features, projected magnetic field strengths and X-ray emission levels, with the different concentrations having separations of order 90$^{\circ}$. This configuration is
frequently found together with a large polar spot, and
appears characteristic of this type of active cool dwarf star (Hussain et al., 2007).
\section{Detailed Data Analysis}
\subsection{Radio Spectral indices}
\label{indices}
The spectral indices between 2.37 and 4.80~GHz and between 4.80 and 8.64~GHz are listed in separate columns in Tables 1 and 2. These are computed assuming a power-law spectrum as defined in Section 2.2.1.
The indices in Table~\ref{table:statistics} are those obtained from the averaged flux densities for each observing session's data, including both the background and flare emission. The error assigned to each spectral index and their differences is that obtained by propagating the bracketed errors assigned to the appropriate flux densities and spectral indices (cf.\ Bevington, 1969). The final column provides a test of the significance of the differences between the pairs of spectral indices. For AB~Dor the spectral indices appear to change from positive to negative near a frequency of 4.8~GHz, indicating that the frequency at which the radio corona changes from being generally optically thick to thin is not far from this frequency. This effect is also apparent in AB Dor B on two of the three nights, and while unconfirmed for the last run the low S/N of the
data should be kept in mind.
Table~\ref{table:flares} identifies the flares in AB Dor's emission and lists their total durations; in addition, the excess peak flux density above an estimated background emission is given at each frequency. The error assigned to the excess peak flux density is greater than that in Table~\ref{table:statistics} because the difference involves two measures, each of which has its own error. Nevertheless, the spectral index is positive below, and negative above, $\sim$4.80~GHz. The significance of this change is determined from the last column; the relevant quantity being found with an overall significance that ranges between 1.7 and 5.5 sigma. The spectral turnover result can be considered as essentially well-established from the higher frequency ranges, however. The relatively lower significance of results involving the S-band integrations has been discussed above, and does not change the main finding about the spectral index.
Weighted means of $\alpha^{1}$ and $\alpha^{2}$ (weights $\sigma^{-2}$) yield $<\alpha^{1}>$~=~$+$1.02$\pm$0.15 and $<\alpha^{2}>$~=~$-$~0.33$\pm$0.07.
\subsection{Polarization}
\label{polarisation}
We tested the circular polarization fractions for AB Dor A and B listed in Table 1 for 21 Nov 2006 by making contour plots at 4.80 and 8.64 GHz
in Stokes V, using all uv data for that date.
A comparison between the V/I results in Table 1 and the two contour maps shown in Figure~7 shows they are in reasonable agreement.
At 4.80 GHz, Table 1 shows that AB Dor A has a significance level of 1-$\sigma$, consistent with its weak V contours.
Component AB Dor B has much more intense V contours, consistent with the 5-$\sigma$ significance in Table 1.
We then examined the V/I values for the 4.80 GHz data of 08 Jan 2007. The 13 twenty-five min integrations that form its radio light curve
in Figure~6 were sub-divided into 65 five-min integrations in order to search for episodes of ECME,
which can have short durations in late-type stars (Slee et al. 2008).
This analysis yielded two interesting conclusions:
(1) 4 of the 5-min integrations (spread over $\sim$10 h) showed significant polarization fractions for component B with $\sigma$-values
of 3, 4, 7, 9.
(2) None of the polarization fractions for component A had $\sigma$-values exceeding 0.8, and most were much lower.
\begin{figure*}
\begin{center} $
\begin{array} {cc}
\hspace{-0.8cm}
\includegraphics[scale=0.35, angle=-90]{21xv.png} &
\includegraphics[scale=0.35, angle=-90]{21cv.png}
\end{array}$
\caption{Contour maps of AB Dor A and B (northern component) in the Stokes V polarization at 8.64~GHz (left) and 4.80~GHz (right), using all the data from 15 baselines on 2006 November 21 and robust weighting. The highest and lowest contour levels at 8.64~GHz are 2.299 and 0.120~mJy~beam$^{-1}$ respectively. The rms system noise level is 47~$\mu$Jy~beam$^{-1}$ and the FWHM (full width at half maximum) of the restoring beam (lower left) is 3.36$\times$2.84~arcsec, with major axis in PA = 71.2$^{\circ}$. At 4.80~GHz, the maximum and minimum contour levels are 0.423 and 0.086~mJy~beam$^{-1}$ respectively,
and the rms system noise level is 39~$\mu$Jy~beam$^{-1}$. The FWHM of the restoring beam
is 4.45$\times$3.51 arcsec, with major axis in PA 72.0$^{\circ}$.}
\label{figure_7}
\end{center}
\end{figure*}
There is an indication that the frequency dependence in spectral index of AB Dor B changes sign from
2006 November 21 to 2007 January 08. In that regard it differs from the more
consistent spectral behaviour of AB Dor A.
It would be useful to try to check the behaviour of AB Dor B
with VLBI observations.
\subsection{Flares}
\label{flare_power}
In order to estimate the radio power emitted during the flares listed in Table~\ref{table:flares} we need a spectral model.
For this purpose, we used the 5-baseline-data at 2.37, 4.80, and also 8.64 GHz, if a thoroughly calibrated contour plot showed an image well above the system noise and defects produced by incomplete sampling of the $uv$-plane. These high angular resolution data provided accurate off-sets from the phase centre for AB Dor A and B. At 2.37 GHz, the 5 baseline data provided reliable (though noisy) mean values for the flux densities over the two long runs. This gave rise to our two estimates of the spectral indices between 2.37 and 4.80 GHz listed in Tables 1 and 2.
From this, we propose the following spectral model over the frequency range from 10 MHz to 300~GHz:
\begin{itemize}
\item (a) S($\nu$) = A$\nu^{1.02}$ between 0.01 and 4.80~GHz
\item (b) S($\nu$) = B$\nu^{-0.33}$ between 4.80 and 20~GHz
\item (c) S($\nu$) = C$\nu^{-1.5}$ between 20 and 300~GHz
\end{itemize}
where S($\nu$) is the flux density and A and B are constants that can be evaluated by substituting in the appropriate equation a measured value of flux density at a particular frequency. For the highest frequency range we have no data directly available, but we can tailor the observed result for the middle range of frequencies into a feasible high-frequency fall-off, using the known properties of solar flares as a guide (e.g.\ Hurford, 1986).
Using the 4.80~GHz peak flux density in Table~\ref{table:flares} of 5.95$\pm$0.26~mJy in flare No. 2 on 2007 January 08 as the reference, we find A = 1.20, B = 9.99, C = 332. Examination of the errors of the flux measures at the respective frequencies show A can be estimated to within $\sim35$\% of its real value, while B should be accurate to within
$\sim10$\% for a given flare. The high frequency extension illustrates the point that a considerable amount of flare radiation is likely beyond what is typically considered
in microwave flux calculations.
Integrating over the three frequency ranges with this model, we find the peak power emitted by flare 2 on January 08 is 5.12$\times$10$^{18}$~W. Here we use a distance to AB Dor of 14.9~pc. Table~\ref{table:flares} indicates that a typical flare duration to half--brightness is about 3~hours, yielding a total energy output of 5.5$\times$10$^{22}$~J. Since the peak flux densities of the identified flares in Table~\ref{table:flares} vary by about a factor of 5-7 at 4.80 and 8.64~GHz, but their durations appear similar, their total energy outputs are likely to vary by similar proportions.
These estimates of peak and total powers are necessarily uncertain, because we have direct values of the spectral indices only over a small frequency range and, even then, the averaged indices $\alpha^{1}$ and $\alpha^{2}$ have uncertainties of around 15 and 20\% respectively. But the uncertainties at the low and intermediate frequencies have relatively small weight on the total result: it is in the range above 20~GHz that there may be significant departures from the radio flare energies typically considered. Even with the comparatively steep spectral index of --1.5, this region would contribute more than half of the total radio emission.
\subsection{Intensity structure}
\label{ISC}
A good idea of the temporal intensity structure in the radio measurements can be gained by plotting a scatter diagram of the data after subdivision into 5--min.\ integrations. Figure~\ref{AB_Dor_correlations} shows the 8.64 GHz 5-min. integrations plotted against those at 4.80 GHz for the long observing periods on November 21 and January 08. A linear regression line with a slope of 0.72$\pm$0.05 has been fitted to the 153 points in the plot; the correlation coefficient is 0.78.
\begin{figure}
\hspace{-0.5cm}
\includegraphics[scale=1, angle=0]{ABDor_Fig3.png}
\caption{Correlation between 153 five--minute integrations of the AB Dor data at 4.80 and 8.64~GHz on 2006 November 21 and 2007 January 08. The regression line has a slope of 0.72$\pm$0.05 and the correlation co-efficient of 0.78 can
be regarded as highly significant for a 153 point sample. The individual points in the plot are formed from flux densities with rms residuals of $\sim$0.4~mJy.}
\label{AB_Dor_correlations}
\end{figure}
Figure~\ref{AB_Dor_correlations} confirms that most of the intensity variability is in the range 2--7~mJy. The means of
the five minute integrations and their standard deviations at 4.80 and 8.64~GHz are 3.96$\pm$1.43 and 3.30$\pm$1.32 respectively. Typical intensity variability about these levels is thus in the range 2.6--2.8~mJy at the two frequencies. Individual intensities for each point in Figure~8 have measurement errors between 0.35 and 0.45~mJy.
Previous observations of active stars (including AB Dor) have been published by Budding et al.\ (2002), showing that the highest correlation between 4.80 and 8.64~GHz intensity variations occurred in a variety of active systems if the set of observations at one frequency is shifted in time with respect to the other set. In most cases, the cross-correlation functions were symmetrical about a shift consistent with flux variations at 4.80~GHz being later than those at 8.64~GHz. The relatively high cross-correlation peaks found in such studies give a basic confirmation of the underlying reality
of the observed phenomena, as well as, in a more detailed way, allowing physical inferences about the locations and
behaviour of the emerging wavefronts. Similar effects were reported by Osten et al.\ (2002) and discussed
in the solar context by Bastien et al.\ (1998). The timescale of observed time-shifts between correlated events at different frequencies is consistent, in general, with the emission being propagated by magnetohydrodynamic waves through a highly energized, turbulent corona.
In the present experiment, we have used 5--min.\ integrations from the long observational series on 2006 November 21 and 2007 January 08 as shown in Figure~\ref{AB_Dor_correlations}. In order to compute the cross-correlation function, we shift the 8.64~GHz set of 5-min.\ integrations with respect to the fixed 4.80~GHz set, starting at zero shift and going to a maximum of +7 or +8 shifts. At each shift, we fit a linear regression equation and note the correlation coefficient and its significance. The process is then reversed, with the 4.80~GHz set of 5-min.\ integrations being shifted with respect to the fixed 8.64~GHz set to form a negative set of shifts. The auto-correlation functions can be computed in a similar way, although in this case the process need not be reversed. The adopted auto-correlation function is the geometric mean of the 4.80 and 8.64~GHz functions. The cross-correlations are plotted and fitted with Gaussians, from which the shifts listed
in the caption to Fig~9 were determined. These time shifts are significant at the 3-4 $\sigma$ level.
A good description of the applied technique that includes appropriate
discussion of significance tests was given by Moore (1958).
The use of the Gaussian fitting function, apart from its empirical validity in the region of the maximum, can be understood intuitively in terms
of the scattering effect of the turbulent medium on emerging radiation fronts, given their inferred spatial separation.
\begin{figure*}
\includegraphics[width=16.5cm, angle=0]{ABDor_Fig4.png}
\vspace{-0.6cm}
\caption{Correlation between 4.80 and~8.64 GHz intensity variations in two extensive sets of observations; each data-set has been subdivided into 5-minute integrations. The filled circles denote the cross-correlation function, which is smoothed by fitting a Gaussian. The crosses define the averaged 4.80 and 8.64~GHz auto-correlation function. The derived negative shifts of the cross with respect to the auto-correlation functions imply that 4.80~GHz intensity variations follow those at 8.64~GHz by 4.5$\pm$1.3~min on 2006 November 21 and 4.0$\pm$1.1~min on 2007 January 08. The procedure is described in Section~\ref{ISC} }.
\label{AB_Dor_correlations_48_86}
\end{figure*}
The resulting correlation functions are presented in Figure~\ref{AB_Dor_correlations_48_86}, which shows that the cross-correlations are systematically displaced in a negative direction with respect to the mean auto-correlations.
The trend of negative shifts for the Nov 21, 2006 data yielded a mean lag of --4.50$\pm$1.30 min.
The more regular cross-correlation function for the Jan 08, 2007 observations gave a mean shift of --4.00$\pm$1.13 min.
The lags should thus be regarded as confirmed at the 3$\sigma$ level.
Similar findings were reported by Budding et al.\ (2002), whose correlation analysis of four long observations of AB Dor in 1994 and 1997 yielded shifts of 4.0 to 6.5~min. on three of the four nights (Table 1 in Budding et al., 2002). Typical errors
for these determinations were given as $\sim2$ min, so while differences in the X to C-band shifts for individual
flares cannot be reliably specified, the shifts themselves can be regarded as well-confirmed with closely comparable results
obtained on 5 out of 6 independent data-sets.
Apart from these intercorrelations of the microwave data-sets themselves, there is interest in the correlation that such data may have with contemporaneous observations at higher photon energies (Forbrich et al., 2011). Such correlations can illuminate physical understanding of flare energetics. Figure~3 in Forbrich et al.\ (2011) shows detailed plots of the microwave variations against contemporaneous X-ray observations for 5 well-known active stars, including the physically
comparable object CC Eri, mentioned above. In fact, if we use the measured flare fluxes: in the X-ray region from Section 2.1, and in the C-band from Table 2, and convert to the same (cgs) luminosity scales used in Forbrich et al.'s Figure~3, we will find representative points for these AB Dor flares (X-ray: 1.1 - 1.6$\times$10$^{30}$;
C-band: 5 - 10$\times$10$^{14}$ Hz$^{-1}$) reasonably typical among the accumulation
of points for CC Eri towards the lower left of the diagram (see Figure~\ref{AB_Dor_xradcor}).
\begin{figure}
\vspace{-0.3cm}
\includegraphics[width=7.2cm, angle=-90]{ccxrad2.png}
\caption{ A scanned version of part of Figure~3 from Forbrich et al.\ (2011), covering
observed microwave and X-ray log(flux) combinations for CC Eri flares, is here shown with
a cross at 14.9, 30.12 to represent range of log (flux) values in X-rays and C-band for the flares of AB Dor
discussed herein. It would appear that our results for AB Dor are well within the
general correlation discussed by Forbrich et al., and while comparable to those of CC Eri
are somewhat to the faint side of the distribution for this class of object.}
\label{AB_Dor_xradcor}
\end{figure}
The existence of such correlations between X-ray fluxes and microwave emission is taken, in a general way, to support the view that the radio emission originates from the tail of the Maxwellian electron distribution of a very hot plasma through the gyrosynchrotron process; the same, flare-generating, electron population being involved in both emissions. Although this model has the appeal of physical unification, detailed matching of the quantities involved leads to
complications associated with the structure, properties and geometry
of the sources. These are not the same for all events, as discussed below.
\section{Discussion}
\begin{figure*}
\hspace{0cm}
\includegraphics[width=12cm,angle=-90.]{fig9n2.png}
\caption{Superposition of the microwave (4.8 GHz) and X-ray light curves for AB Dor.
The vertical bars on the microwave points indicate measurement uncertainties, while
the horizontal ones show the duration of the corresponding integrations.}
\label{AB_Dor_xfigt}
\end{figure*}
Budding et al.\ (2009) reviewed in some detail photometry and spectroscopy of AB Dor that formed a subset of the present multiwavelength study.
The broadband light curves were dominated by one outstanding macula, while the spectroscopic features suggested several chromospheric activity sites, one strong concentration of which, located near phase zero, could be associated with the main macula. These findings led to the surmise of a large bipolar surface structure that was linked to the site of contemporaneous multiwavelength activity. Our subsequent reduction and analysis of the Suzaku data are compromised, to some extent, by the lack of a clear resolution between the two
sources. This certainly holds for the quiescent emission levels, where previous long duration
observations by the Chandra satellite point to the source AB Dor A predominating by a factor of
typically $\sim$6. Moreover, our present X-ray data
show clear links between the major X-ray and microwave flares, which support the idea of one large magnetodynamic activity generator that may be given an approximate position on the surface of AB Dor during late 2006 and early 2007.
The X-ray data, newly presented in this article,
may have a special role in placing the sequence of
events in flare energization. In this connection,
X-ray spectral indices for the Suzaku data for Nov 2006 and Jan 2007 were measured and their behaviour shown in Figure~4.
The data cover quiescent and flaring phases. Of special interest are the near peak regions, particularly the second flare of the Jan 2007 data that shows an initial impulse and subsequent second peak occurring after the
intermediate microwave peak (see Figure~\ref{AB_Dor_xfigt}).
The higher-energy quiescent fluxes (not reliably measurable in the highest frequency range for quiescent GTIs) were consistent with the full X-ray spectra shown in Figure~2,
and the formula $F_\nu \sim A.\nu^{\delta}$. The spectral index $\delta$ = --3 is often adopted in theoretical
treatments and our X-ray spectral indices are broadly in agreement with this.
Dulk (1985) shows the variation of absorption and emission coefficients for microwave radiation to be quite sensitive
to the value of $\delta$ for gyrosynchrotron emission. The effective temperature has a moderate dependence, while the
degree of circular polarization is less affected. It would clearly be useful to take into account observational evidence
on $\delta$ in more detailed modelling of the radio emission.
An interesting point arising from the juxtaposition of the X-ray and microwave
flux variations (Fig.~\ref{AB_Dor_xfigt})
is whether the initial sharp impulse on
the second flare of Jan 8 (GTI 12) corresponds to
a high-energy trigger preceding the microwave maxima.
This behaviour has been seen in previous similar event sequences
(Budding et al., 2006).
If we compare the spectral indices for GTIs 9, 12 and 13 (Jan 8/9) there is not
much difference between those of 9 and 12 ---
i.e.\ the initial impulse of the second (smaller) flare is no harder
than the decline phase of the first flare. That points to
a much more powerful initial impulse to that first flare
that was not detected during the observations.
All these GTIs do, however, have significantly harder
spectral indices than the quiescent spectrum.
The index hardens to around 0.5 greater than during quiescence, maximizing,
in these data sets, at the
earliest stage of the second flare.
On the other hand, that second flare of Jan 8 is
definitely softer after the intervening microwave peak.
The X-ray spectral index declines by about 30\% in the second X-ray maximum,
which can be associated with near-footpoint bremsstrahlung radiation
(Bastian et al., 1998).
The gentler downward slope of the X-rays
in the post-microwave maxima phases of this flare concurs with the Neupert effect here.
A reasonably self-consistent overall picture for the observed flare sequence
can then be presented, on the basis that there was a first
very large but unseen impulse to the first flare, while for the second
(weaker) flare we observe the whole sequence: initial impulse,
gyrosynchrotron cooling and then bremsstrahlung from precipitated electrons.
This picture can be placed alongside the interpretation of
optical and radio data given by Budding et al.\ (2009), where it was
found reasonable to associate the radio observations of 2006 November and 2007 January to one large and predominating visual macula of around 14-15 deg angular radius, longitude 5-10 deg and latitude around 50-60 deg. A large
chromospheric facular region could also be linked to this macula.
This active region comes into full view near phase zero, when the first of the Jan 2007
flares occurs. The second major flare of the Jan 2007 observations (as well as the Nov 2006 one), however, occurs around phases 0.6-0.7, which is about the location of `facula b', the second major chromospheric feature
detected from the H$_{\alpha}$ emission line structure by Budding et al.\ (2009).
Although about the same, or slightly weaker, in X-rays (the first flare's
X-ray start looks to have been missed), this flare has a stronger microwave
peak. Although this flare may then be of intrinsically lower energy
than the first one, a reasonable inference is that the source geometry happens
to give us a better view, including its initial
impulsive phase.
It would be unwise to put too much weight on this scenario from just the limited amount of data presented here -- but our interpretation
is worth retaining for further testing against more data that should be
expected in future.
\section*{Acknowledgments}
We thank the time allocation committee of the Australia Telescope National Facility for two generous allocations on the ATCA to support this multiwavelength campaign. The original proposal and arrangements for acquiring time on the Suzaku satellite actually came from Dr Mark Audard, of Geneva; to whom we express
our gratitude for his early inputs and suggestions. Dr Grigor Tsarevsky of the Sternberg Institute, Moscow also contributed greatly with advice and observational assistance.
We thank also the staff of the Armagh Observatory, who helped N.E.\
progress this work, particularly Dr G.\ Ramsay in relation to the
processing of the Suzaku data and subsequent comments on
its discussion. N.E.'s study leave in Armagh was
supported by the EU Erasmus Programme.
Detailed comments from a reviewer have led to
a much more substantiated presentation.
|
\section{Introduction}\label{sect1}
\renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0}
\setcounter{thm}{0} \setcounter{lemma}{0}
In this work we intend to deal with Bahadur-Kiefer type processes that
are based on partial sums and their renewals of weakly, as well
as strongly, dependent sequences of random variables. In order to initiate
our approach, let $\{Y_0,Y_1,Y_2,\ldots\}$ be random variables which have the
same marginal distribution and, to begin with, satisfy the following
assumptions:
\begin{itemize}
\item[{\rm (i)}]\ ${\rm E} Y_0=\mu>0$;
\item[{\rm (ii)}]\ ${\rm E}(Y_0^{2})<\infty$.
\end{itemize}
In terms of the generic sequence $\{Y_j,\, j=0,1,2,\ldots\}$, with
$t\geq 0$, we define
\begin{eqnarray}
S(t)&:=&\sum_{i=1}^{[t]}Y_i,\label{st}\\
N(t)&:=&\inf\{s\geq 1:\, S(s)>t\},\label{nt}\\
Q(t)&:=& S(t)+\mu N(\mu t)-2\mu t,\label{qt}
\end{eqnarray}
whose respective appropriately normalized versions will be used in studying
partial sums, their renewals, Bahadur-Kiefer type processes when the random
variables in the sequence $Y_i,\, i=0,1,\ldots$ are weakly or strongly
dependent.
The research area of what has become known as Bahadur-Kiefer processes was
initiated by Bahadur \cite{Bahadur1966} (cf. also Kiefer \cite{Kiefer1967})
who established an almost sure representation of i.i.d. random variables
based sample quantiles in terms of their empiricals. Kiefer
\cite{Kiefer1970} substantiated this work via studying the deviations between
the sample quantile and its empirical processes. These three seminal papers
have since been followed by many related further investigations (cf., e.g.,
Cs\"org\H o and R\'ev\'esz \cite{CsR1978}, [12, Chapter 5],
Shorack \cite{Shorack}, Cs\"org\H o \cite{Cs1983}, Deheuvels and Mason
\cite{DM}, \cite{DM2}, Deheuvels \cite{D921}, \cite{D922}, Cs\"org\H o and
Horv\'ath [7, Chapters 3-6], Cs\"org\H o and Szyszkowicz \cite{CsSz},
and references in these works).
It follows from the results of Kiefer \cite{Kiefer1970}, and also from
Vervaat \cite{Verv1}, \cite{Verv2} as spelled out in Cs\'aki {\it et al.}
\cite{CsCsRS}, that the original i.i.d. based Bahadur-Kiefer process
cannot converge weakly to any non-degenerate random element of the
$D[0,1]$ function space. On the other hand, Cs\"org\H o {\it et al.}
\cite{CsSzW} showed the opposite conclusion to hold true for long-range
dependence based Bahadur-Kiefer processes. For an illustration and discussion
of this conclusion, we refer to the Introduction and Corollary 1.2 of Cs\'aki
{\it et al.} \cite{CsCsK}. For further results along these lines, we refer to
Cs\"org\H o and Kulik \cite{CsK1}, \cite{CsK2}.
The study of the almost sure asymptotic behaviour of Bahadur-Kiefer type
processes for sums and their renewals in the i.i.d. case was initiated by
Horv\'ath \cite{Horv}, Deheuvels and Mason \cite{DM}, and augmented by
further references and results as in Cs\"org\H o and Horv\'ath [7,
Chapter 2].
Vervaat \cite{Verv1}, \cite{Verv2} initiated the study of limit theorems in
general for processes with a positive drift and their inverses. For results
on the asymptotic behaviour of integrals of Bahadur-Kiefer type processes
for sums and their renewals, the so-called Vervaat processes, we refer to
\cite{CsCsRS} in the i.i.d. case, \cite{CsakiCsorgoKulik} in the weakly
dependent case, and \cite{CsCsK} in the strongly dependent case.
Back to the topics of this paper on Bahadur-Kiefer type processes for sums
and their renewals, the forthcoming Section 2 is concerned with
the weakly dependent case, and Section 3 concludes results in terms of
long-range dependent sequences of random variables. Both of these sections
contain the relevant proofs as well.
\section{Weakly dependent case}
\renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0}
In this section we deal with weakly dependent random variables based
Bahadur-Kiefer type processes. First we summarize the
main results in the case when $Y_i$ are i.i.d. random variables with finite
4-th moment.
\medskip\noindent
{\bf Theorem A} {\it Assume that $\{Y_i,\, i=0,1,\ldots\}$ are i.i.d. random
variables with ${\rm E} Y_0=\mu>0$, ${\rm E} (Y_0-\mu)^2=\sigma^2>0$, and
${\rm E} Y_0^4<\infty$. Then we have
\begin{equation}
Q(T)=\sigma
\left(W(T)-W\left(T-\frac{\sigma}{\mu}W(T)\right)\right)+o_{{\rm a.s.}}(T^{1/4}),
\quad {\rm as}\ \ T\to\infty,
\label{qtinv}
\end{equation}
\begin{equation}
\limsup_{T\to\infty}\frac{\sup_{0\leq t\leq T}|Q(t/\mu)|}
{(T\log\log T)^{1/4}(\log T)^{1/2}}=\frac{2^{1/4}\sigma^{3/2}}{\mu^{3/4}},
\quad {\rm a.s.},
\label{qtlimsup}
\end{equation}
\begin{equation}
\lim_{T\to\infty}\frac{\sup_{0\leq t\leq T}|Q(t/\mu)|}
{(\log T)^{1/2}(\sup_{0\leq t\leq T}|\mu N(t)-t|)^{1/2}}=
\frac{\sigma}{\mu^{1/2}},
\quad {\rm a.s.},
\label{qtlim}
\end{equation}
\begin{equation}
\lim_{T\to\infty}{\rm P}(T^{-1/4}|Q(T/\mu)|\leq y)=
2\int_{-\infty}^\infty \Phi(y\mu^{3/4}\sigma^{-3/2}|x|^{-1/2})
\varphi(x)\, dx-1,
\label{qtl}
\end{equation}
where $\Phi$ is the standard normal distribution function and $\varphi$ is
its density.
}
\medskip
We note that (\ref{qtinv}) and (\ref{qtl}) are due to Cs\"org\H o and
Horv\'ath \cite{CsH}, (\ref{qtlimsup}) is due to Horv\'ath \cite{Horv} and
(\ref{qtlim}) is due to Deheuvels and Mason \cite{DM}. All these results
can be found in \cite{CsH}.
For the case of i.i.d. random variables when the 4-th moment does not exist,
we refer to Deheuvels and Steinebach \cite{DS}.
In this section we assume that $S(t)$ can be approximated by a standard
Wiener process as follows.
\noindent
{\bf Assumption A} {\it On the same probability space there exist a sequence
$\{Y_i,\ i=0,1,2,\ldots\}$ of random variables, with the same marginal
distribution, satisfying assumptions {\rm (i)} and {\rm (ii)}, and a
standard Wiener process $W(t),\ t\geq 0$, such that
\begin{equation}
\sup_{0\leq t\leq T}|S(t)-\mu t-\sigma W(t)|=O_{{\rm a.s.}}(T^\beta)
\label{appr}
\end{equation}
almost surely, as $T\to\infty$, with $\sigma>0$,
where $S(t)$ is defined by {\rm (\ref{st})} and $\beta<1/4$.}
In the case of $1/4\leq \beta<1/2$, there is a huge literature on strong
approximation of the form (\ref{appr}) for weakly dependent random variables
$\{Y_i\}$. The case $\beta<1/4$ is treated in Berkes {\it et al.}
\cite{BerkesLiuWu}, where Koml\'os-Major-Tusn\'ady
\cite{KomlosMajorTusnady1975} type strong approximations as in
(\ref{appr}) are proved under fairly general assumptions of dependence.
For exact statements of, and conditions for, strong approximations that
yield (\ref{appr}) to hold true for the partial sums as in Assumption A,
we refer to \cite{BerkesLiuWu}.
\begin{thm}
Under {\rm Assumption A} all the results {\rm (\ref{qtinv}), (\ref{qtlimsup}),
(\ref{qtlim}) and (\ref{qtl})} in {\rm Theorem A} remain true.
\end{thm}
\noindent
{\bf Proof}. In fact, we only have to prove (\ref{qtinv}), for the other
results follow from the latter. It follows from \cite{CsH}, Theorem 1.3
on p. 37, that under Assumption A we have
$$
\limsup_{T\to\infty}\frac{\sup_{0\leq t\leq T}
\left|\frac{t}{\mu}-N(t)-\frac{\sigma}{\mu}W(t/\mu)\right|}
{(T\log\log T)^{1/4}(\log T)^{1/2}}
=2^{1/4}\sigma^{3/2}\mu^{-7/4}\quad {\rm a.s.}
$$
and also
$$
\sup_{0\leq t\leq T}|\mu t-\mu S(N(\mu t))|=O_{{\rm a.s.}}(T^\beta)
$$
as $T\to\infty$. Hence, as $T\to\infty$, we arrive at
$$
Q(T)=S(T)+\mu N(\mu T)-2\mu T=S(T)-\mu T-\left(S(N(\mu T))-\mu N(\mu T)\right)
+O_{{\rm a.s.}}(T^\beta)
$$
$$
=\sigma(W(t)-W(N(\mu T)))+O_{{\rm a.s.}}(T^\beta)=
$$
$$
\sigma
\left(W(T)-W\left(T-\frac{\sigma}{\mu}W(T)\right)\right)+o_{{\rm a.s.}}(T^{1/4}),
$$
i.e., having (\ref{qtinv}) as desired.
\qed
\section{Strongly dependent case}
\renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0}
\setcounter{thm}{0} \setcounter{lemma}{0}
In this section we deal with long range (strongly) dependent sequences,
based on moving averages as defined by
\begin{equation}
\eta_j=\sum_{k=0}^\infty \psi_k \xi_{j-k},\quad j=0,1,2,\ldots,
\label{linear}
\end{equation}
where $\{\xi_k,\, -\infty<k<\infty\}$ is a double sequence of independent
standard normal random variables, and the sequence of weights
$\{\psi_k, \, k=0,1,2,\ldots\}$ is square summable. Then
${\rm E}(\eta_0)=0$, ${\rm E}(\eta_0^2)=\sum_{k=0}^\infty\psi_k^2=:
\sigma^2$ and, on putting $\tilde\eta_j=\eta_j/\sigma$, $\{\tilde\eta_j,\,
j=0,1,2,\ldots\}$ is a stationary Gaussian sequence with
${\rm E}(\tilde\eta_0)=0$ and ${\rm E}(\tilde\eta_0^2)=1$. If
$\psi_k\sim k^{-(1+\alpha)/2}\ell(k)$ with a slowly varying function,
$\ell(k)$, at infinity, then
${\rm E}(\eta_j\eta_{j+n})\sim b_\alpha n^{-\alpha}\ell^2(n)$,
where the constant $b_\alpha$ is defined by
$$
b_\alpha=\int_0^\infty x^{-(1+\alpha)/2}(1+x)^{-(1+\alpha)/2}\, dx.
$$
Now let $G(\cdot)$ be a real valued Borel measurable function,
and define the subordinated sequence $Y_j=G(\tilde\eta_j)$, $j=0,1,2,\ldots$.
We assume throughout that $J_1:={\rm E}(G(\tilde\eta_0)\tilde\eta_0)\not=0$.
We say in this case that the Hermite rank of the function $G(\cdot)$ is
equal to 1 (cf. Introduction of \cite{CsCsK}).
For $1/2<H<1$ let $\{W_H(t),\, t\geq 0\}$ be a fractional Brownian
motion (fbm), i.e., a mean-zero Gaussian process with covariance
\begin{equation}
{\rm E} W_H(s)W_H(t)=\frac{1}{2}(s^{2H}+t^{2H}-|s-t|^{2H}).
\label{covar}
\end{equation}
Based on a strong approximation result of Wang {\it et al.}
\cite{WangLinGulati2003}, what follows next, was proved in Section 2 of
Cs\'aki {\it et al.} \cite{CsCsK}.
\medskip\noindent
{\bf Theorem B}
{\it Let $\eta_j$ be defined by {\rm (\ref{linear})} with $\psi_k\sim
k^{-(1+\alpha)/2}$, $0<\alpha<1$, and put $\tilde\eta_j=\eta_j/\sigma$
with $\sigma^2:={\rm E}(\eta_0^2)=\sum_{k=0}^\infty \psi_k^2$. Let $G(\cdot)$
be a function whose Hermite rank is $1$, and put $Y_j=G(\tilde\eta_j)$,
$j=0,1,2,\ldots$. Furthermore, let $\{S(t),\, t\geq 0\}$
be as in {\rm (\ref{st})} and assume condition {\rm (ii)}. Then, on
an appropriate probability space for the sequence
$\{Y_j=G(\tilde\eta_j),\, j=0,1,\ldots\}$, one can construct a fractional
Brownian motion $W_{1-\alpha/2}(\cdot)$ such that, as $T\to\infty$, we have
\begin{equation}\label{eq:3/1}
\sup_{0\le t\le T}\left|S(t)-\mu t-
\frac{J_1 \kappa_{\alpha}}{\sigma}W_{1-\alpha/2}(t)\right|=
o_{{\rm a.s.}}(T^{\gamma/2+\delta}),
\end{equation}
where $\mu={\rm E}(Y_0)$,
\begin{equation}
\kappa_{\alpha}^2=
2\frac{\int_0^{\infty}x^{-(\alpha+1)/2}(1+x)^{-(\alpha+1)/2}\,
dx}{(1-\alpha)(2-\alpha)},
\label{kalpha}
\end{equation}
$\gamma=2-2\alpha$ for $\alpha<1/2$,
$\gamma=1$ for $\alpha\geq 1/2$ and $\delta>0$ is arbitrary}.
{\it Moreover, if we also assume condition {\rm (i)}, then, as $T\to\infty$, }
\begin{equation}
\sup_{0\leq t\leq T}\left |\mu N(\mu t)-\mu t+\frac{J_1
\kappa_\alpha}{\sigma}
W_{1-\alpha/2}(t)\right
|=o_{{\rm a.s.}}(T^{\gamma/2+\delta}+T^{(1-\alpha/2)^2+\delta}),
\label{eq:main2}
\end{equation}
{\it with $\gamma$ as right above, and arbitrary $\delta>0$.}
\medskip
Now, for use in the sequel, we state iterated logarithm results for
fractional Brownian motion and its increments, which follows from
Ortega's extension in \cite{Ortega1984} of Cs\"org\H o and R\'ev\'esz
\cite{CsR1979}, [12, Section 1.2].
\medskip\noindent
{\bf Theorem C} {\it
For $T>0$ let $a_T$ be a nondecreasing function of $T$ such that
$0<a_T\leq T$ and $a_T/T$ is nonincreasing. Then
\begin{equation}
\limsup_{T\to\infty}\frac{\sup_{0\leq t\leq T-a_T}\sup_{0\leq s\leq a_T}
|W_{1-\alpha/2}(t+s)-W_{1-\alpha/2}(t)|}
{a_T^{1-\alpha/2}(2(\log T/a_T+\log\log T))^{1/2}}=1\quad {\rm a.s.}
\label{Ortega}
\end{equation}
If $\lim_{T\to\infty}(\log(T/a_T))/(\log\log T)=\infty$, then we have $\lim$
instead of $\limsup$ in {\rm (\ref{Ortega})}.}
\medskip
First we give an invariance principle for $Q(T)$ defined by (1.3) if
$\gamma/2<(1-\alpha/2)^2$, which corresponds to the i.i.d. case when the
forth moment exists. Equivalently, we assume that
\begin{equation}
0<\alpha<2-\sqrt{2}.
\label{alfa}
\end{equation}
Note that in (\ref{qtinv2}) below, the random time argument of
$W_{1-\alpha/2}$ is strictly positive for large enough $T$ with probability 1.
So, without loss of generality, we may define
$W_{1-\alpha/2}(T-u)=0$ if $u>T$.
\begin{thm} Under the conditions of {\rm Theorem B}, including {\rm (i)}
and {\rm (ii)}, assuming {\rm (\ref{alfa})}, as $T\to\infty$, we have
$$
Q(T)=
\frac{J_1\kappa_\alpha}{\sigma}(W_{1-\alpha/2}(T)-W_{1-\alpha/2}(N(\mu T))
+o_{{\rm a.s.}}(T^{\gamma/2+\delta})
$$
\begin{equation}
=\frac{J_1\kappa_\alpha}{\sigma}
\left(W_{1-\alpha/2}(T)-W_{1-\alpha/2}\left(T-
\frac{J_1\kappa_\alpha}{\sigma\mu}W_{1-\alpha/2}(T)\right)\right)
+o_{{\rm a.s.}}(T^{\gamma/2+\delta}).
\label{qtinv2}
\end{equation}
\end{thm}
\noindent {\bf Proof.} Put $c=J_1\kappa_\alpha/\sigma$. Then
$$
Q(T)=S(T)-\mu T+\mu N(\mu T)-\mu T
$$
$$
=cW_{1-\alpha/2}(T)+
o_{{\rm a.s.}}(T^{\gamma/2+\delta})+\mu (N(\mu T)-T).
$$
But
$$
\mu(T-N(\mu T))=S(N(\mu T))-\mu N(\mu T)+\mu T-S(N(\mu T))
$$
$$
=cW_{1-\alpha/2}(N(\mu T))+o_{{\rm a.s.}}((N(\mu T))^{\gamma/2+\delta})+
\mu T-S(N(\mu T)),
$$
and using (\ref{eq:main2}) and Theorem C, we have
$$
cW_{1-\alpha/2}(N(\mu T))=cW_{1-\alpha/2}\left(T-
\frac{c}{\mu}W_{1-\alpha/2}(T)+o_{{\rm a.s.}}(T^{\gamma/2+\delta}
+T^{(1-\alpha/2)^2+\delta})\right)
$$
$$
=cW_{1-\alpha/2}\left(T-\frac{c}{\mu}W_{1-\alpha/2}(T)\right)+
o_{{\rm a.s.}}(T^{(\gamma/2+\delta)(1-\alpha/2)}+T^{(1-\alpha/2)^3}).
$$
On the other hand (cf. \cite{CsCsK}), $N(\mu T)=O_{{\rm a.s.}}(T)$ and
$$
\mu T-S(N(\mu T))=o_{{\rm a.s.}}(T^{\gamma/2+\delta}).
$$
Since $(1-\alpha/2)^3\leq \gamma/2< (1-\alpha/2)^2$, this
dominates all the other remainder terms
in the proof. Thus the proof of Theorem 3.1 is now complete.
\qed
The proof of Theorem 3.1 also yields the following result.
\begin{proposition}
As $T\to\infty$,
$$
\mu T-\mu N(\mu T))=\frac{J_1\kappa_\alpha}{\sigma}
W_{1-\alpha/2}\left(T-\frac{J_1\kappa_\alpha}{\sigma\mu}W_{1-\alpha/2}(T)
\right)+o_{{\rm a.s.}}(T^{\gamma/2+\delta}).
$$
\end{proposition}
Now we are to give a limsup result for $Q(\cdot)$. For this we need a
Strassen-type functional law of the iterated logarithm for fbm, due to
Goodman and Kuelbs \cite{GK1991}.
\newpage
\medskip\noindent{\bf Theorem D} {\it Let
$$
{\bf K}=\{T_H g(t),\, 0\leq t\leq 1,\,
\int_{-\infty}^1 g^2(u)\, du\leq 1\},
$$
where
$$
T_H g(t)=\frac{1}{k_H}\int_0^t (t-u)^{H-1/2}g(u)\, du
+\frac{1}{k_H}\int_{-\infty}^0(t-u)^{H-1/2}-(-u)^{H-1/2})g(u)\, du,
$$
and
$$
k_H^2=\int_{-\infty}^0((1-s)^{H-1/2}-(-s)^{H-1/2})^2\, ds
+\int_0^1(1-s)^{2H-1}\, ds.
$$
Then, almost surely, ${\bf K}$ is the set of limit points of the net of
stochastic processes
\begin{equation}
\frac{W_H(nt)}{(2n^{2H}\log\log n)^{1/2}},\, 0\leq t\leq 1,
\label{flil}
\end{equation}
as $n\to\infty$.}
\begin{thm} Under the conditions of {\rm Theorem 3.1}, we have
\begin{equation}
\limsup_{T\to\infty}
\frac{|Q(T)|}{T^{(1-\alpha/2)^2}(\log\log T)^{1/2-\alpha/4}(\log T)^{1/2}}
=\frac{2^{1-\alpha/4} (J_1\kappa_\alpha)^{2-\alpha/2}}
{\sigma^{2-\alpha/2}\mu^{1-\alpha/2}}\qquad {\rm a.s.}
\label{qtlimsup2}
\end{equation}
\end{thm}
\noindent
{\bf Proof}. It follows from Theorem C that
$$
|W_{1-\alpha/2}(T)|\leq (1+\delta)T^{1-\alpha/2}(2\log\log T)^{1/2}
$$
with probability 1 for any $\delta>0$ if $T$ is large enough. Hence, applying
Theorem C with $a_T=(1+\delta)c/\mu T^{1-\alpha/2}(2\log\log T)^{1/2}$,
$c=J_1\kappa_\alpha/\sigma$, we obtain
$$
c\sup_{|s|\leq a_T}|W_{1-\alpha/2}(T)-W_{1-\alpha/2}(T-s)|\leq
c(1+\delta)a_T^{1-\alpha/2}(2\log T)^{1/2},
$$
almost surely for large enough T. Since $\delta>0$ is arbitrary,
we obtain the upper bound in (\ref{qtlimsup2}).
To obtain the lower bound, we follow the proof in the i.i.d. case, given in
Cs\"org\H o and Horv\'ath \cite{CsH}. On choosing
\begin{displaymath}
g(s)=\left\{
\begin{array}{ll}
\frac{1}{k_H}((1-s)^{H-1/2}-(-s)^{H-1/2}),
&\quad s\leq 0,\\
\frac{1}{k_H}(1-s)^{H-1/2}, &\quad 0<s\leq 1,
\end{array}
\right.
\end{displaymath}
in Theorem D, we have
$$
f(t)=\frac{1}{k_H}\int_{-\infty}^0((t-s)^{H-1/2}-(-s)^{H-1/2})g(s)\, ds
+\frac{1}{k_H}\int_0^t(t-s)^{H-1/2}g(s)\, ds.
$$
It can be seen that $\int_{-\infty}^1 g^2(s)\, ds=1$, and
$\{f(t),\, \, 0\leq t\leq 1\}$ is a continuous increasing function with
$f(0)=0$, $f(1)=1$, and hence by Theorem D it is in ${\bf K}$.
For $0<\delta<1$, on considering the function
\begin{displaymath}
g_\delta(s)=\left\{
\begin{array}{ll}
g(s), &\quad 0\leq s\leq 1-\delta,\\
0, &\quad 1-\delta\leq s\leq 1,
\end{array}
\right.
\end{displaymath}
we define
\begin{displaymath}
f_\delta(t)=\left\{
\begin{array}{ll}
f(t), &\quad 0\leq t\leq 1-\delta,\\
f(1-\delta), &\quad 1-\delta\leq t\leq 1.
\end{array}
\right.
\end{displaymath}
Then it can be seen that the latter function is in ${\bf K}$, and hence it
is a limit function of the net of stochastic processes as in (\ref{flil}).
It follows that there is a sequence $T_k$ of random variables such that,
in our context,
$$
\lim_{k\to\infty}\sup_{0\leq t\leq 1}
\left|\frac{W_{1-\alpha/2}(T_kt)}
{T_k^{1-\alpha/2}(2\log\log T_k)^{1/2}}-f_\delta(t)\right|=0.
$$
Using Theorem C with $a_T=f(1-\delta)c/\mu T^{1-\alpha/2}(2\log\log T)^{1/2}$,
we get
$$
\lim_{T\to\infty}\frac{\sup_{T(1-\delta)\leq t\leq T}
c|W_{1-\alpha/2}(t+a_T)-W(t)|}{ca_T^{1-\alpha/2}(2\log T)^{1/2}}=1
\qquad {\rm a.s.}
$$
Since $\delta$ is arbitrary, and
$\lim_{\delta\to 0}f_\delta(t)=f(t)$, the lower bound
follows as in Cs\"org\H o and Horv\'ath \cite{CsH}, p. 28. This completes the
proof of Theorem 3.2.
\qed
\medskip
Next we give the limiting distribution of $Q(T)$.
\begin{thm} Under the conditions of {\rm Theorem 3.1}, we have
\begin{equation}
\lim_{T\to\infty}{\rm P}\left(Q(T)T^{-(1-\alpha/2)^2}\leq y\right)=
\int_{-\infty}^{\infty}\varphi(x)
\Phi\left(\frac{y \sigma^{2-\alpha/2}\mu^{1-\alpha/2}}
{|x|^{1-\alpha/2}(J_1\kappa_\alpha)^{2-\alpha/2}}\right)\, dx.
\label{limit}
\end{equation}
\end{thm}
\noindent{\bf Proof.}
According to Theorem 3.1 we have to determine the limiting distribution of
$$
c\left(W_{1-\alpha/2}(T)-W_{1-\alpha/2}\left(T-\frac{c}{\mu}
W_{1-\alpha/2}(T)\right)\right),
$$
where $c=J_1\kappa_\alpha/\sigma$. Via the scaling property of fbm, i.e.,
$$
\widetilde W(v):=T^{-1+\alpha/2}W_{1-\alpha/2}(Tv), \quad v\geq 0,
$$
is also an fbm with parameter $1-\alpha/2$. So we have to determine the
limiting distribution of
$$c\left(\widetilde W(1)-\widetilde
W(1-c_1T^{-\alpha/2}\widetilde W(1))\right),$$
as $T\to\infty$, where $c_1=J_1\kappa_\alpha/(\sigma\mu)$.
For $u>0$, the joint distribution of $\widetilde W(1)$, $\widetilde W(u)$
is bivariate normal with density
$$
\frac{1}{2\pi\sigma_1\sigma_2\sqrt{1-r^2}}
\exp\left(-\frac{1}{2(1-r^2)}
\left(\frac{x^2}{\sigma_1^2}-2r\frac{xy}{\sigma_1\sigma_2}+
\frac{y^2}{\sigma_2^2}\right)\right),
$$
where $\sigma_1^2={\rm E}(W_{1-\alpha/2}^2(1))=1$,
$\sigma_2^2={\rm E}(W_{1-\alpha/2}^2(u))=u^{2-\alpha}$ and
$$
r=\frac{1+u^{2-\alpha}-|1-u|^{2-\alpha}}{2\sigma_1\sigma_2}.
$$
Now consider the conditional density
$$
{\rm P}(\widetilde W(u)\in dz|\widetilde W(1)=x)=
\frac{1}{\sigma_2\sqrt{1-r^2}}\varphi\left(\frac{z-r\sigma_2 x}
{\sigma_2\sqrt{1-r^2}}\right)\, dz,
$$
where $u=1-c_1xT^{-\alpha/2}$.
So the density function of $\widetilde W(1)-\widetilde W(u)$ is equal to
$$
{\rm P}(\widetilde W(1)-\widetilde W(u)\in dY)=
\int_{-\infty}^{T^{\alpha/2}/c_1}\frac{1}{\sigma_2\sqrt{1-r^2}}\varphi(x)
\varphi\left(\frac{x-Y-r\sigma_2x}{\sigma_2\sqrt{1-r^2}}\right)\, dx\, dY
$$
and hence its distribution function is
$$
{\rm P}(\widetilde W(1)-\widetilde W(u)\leq Z)=
\int_{-\infty}^{T^{\alpha/2}/c_1}\varphi(x)
\Phi\left(\frac{Z-x+r\sigma_2x}{\sigma_2\sqrt{1-r^2}}\right)\, dx,
\quad -\infty<Z<\infty.
$$
It can be seen that, as $T\to\infty$,
$$
\sigma_2\sqrt{1-r^2}\sim \frac{|c_1x|^{1-\alpha/2}}{T^{\alpha/2-\alpha^2/4}},
$$
$$
\frac{x-xr\sigma_2}{\sigma_2\sqrt{1-r^2}}=O(T^{-\alpha/2+\alpha^2/4}).
$$
Hence, as $T\to\infty$,
$$
{\rm P}(\widetilde W(1)-\widetilde W(u)\leq Z)\sim
\int_{-\infty}^{T^{\alpha/2}/c_1}\varphi(x)
\Phi\left(\frac{ZT^{\alpha/2-\alpha^2/4}}{|c_1x|^{1-\alpha/2}}\right)\, dx.
$$
Putting $Z=yT^{\alpha^2/4-\alpha/2}/c$, and taking the limit $T\to\infty$, we
finally obtain (\ref{limit}).
\qed
\bigskip
\noindent
{\bf Acknowledgement} We wish to thank two referees for their careful
reading of, and constructive remarks on, our manuscript. Research supported
by an NSERC Canada Discovery Grant at Carleton University, Ottawa and by the
Hungarian National Foundation for Scientific Research, No. K108615.
|
\chapter{Introduction}
\label{intro}
In this chapter, we lay out the elements of our novel framework for reinforcement learning \cite{sutton-barto:book}, based on doing temporal difference learning not in the primal space, but in a {\em dual space} defined by a so-called {\em mirror map}. We show how this technical device holds the fundamental key to solving a whole host of unresolved issues in reinforcement learning, from designing stable and reliable off-policy algorithms, to making algorithms achieve safety guarantees, and finally to making them scalable in high dimensions. This new vision of reinforcement learning developed by us over the past few years yields mathematically rigorous solutions to longstanding important questions in the field, which have remained unresolved for almost three decades. We introduce the main concepts in this chapter, from {\em proximal operators} to the {\em mirror descent} and the {\em extragradient method} and its non-Euclidean generalization, the {\em mirror-prox} method. We introduce a powerful decomposition strategy based on {\em operator splitting}, exploiting deep properties of monotone operators in Hilbert spaces. This technical device, as we show later, is fundamental in designing ``true" stochastic gradient methods for reinforcement learning, as it helps to decompose the complex product of terms that occur in recent work on gradient temporal difference learning. We provide examples of the benefits of our framework, showing each of the four key pieces of our solution: the improved performance of our new off-policy temporal difference methods over previous gradient TD methods, like TDC and GTD2 \cite{Sutton09fastgradient-descent}; how we are able to generalize natural gradient actor critic methods using mirror maps, and achieve safety guarantees to control learning in complex robots; and finally, elements of our saddle point reformulation of temporal difference learning. The goal of this chapter is to lay out the sweeping power of our primal dual framework for reinforcement learning. The details of our approach, including technical proofs, algorithms, and experimental validations are relegated to future chapters.
\section{Elements of the Overall Framework}
\subsection{Primal Dual Mirror Maps}
In this section, we provide a succinct explanation of the overall framework, leaving many technical details to future chapters. Central to the proposed framework is the notion of {\em mirror maps}, which facilitates doing temporal learning updates not just in the usual primal space, but also in a dual space. More precisely, $\Phi: {\cal D} \rightarrow \mathbb{R}$ for some domain ${\cal D}$ is a {\em mirror map} if it is strongly convex, differentiable, and the gradient of $\Phi$ has the range $\mathbb{R}^n$ (i.e., takes on all possible vector values). Instead of doing gradient updates in the primal space, we do gradient updates in the dual space, which correspond to:
\[ \nabla \Phi(y) = \nabla \Phi(x) - \alpha \nabla f(x) \]
The step size or learning rate $\alpha$ is a tunable parameter. To get back to the primal space, we use the conjugate mapping $\nabla \Phi^*$, which can be shown to also correspond to the inverse mapping $(\nabla \Phi)^{-1}$, where the conjugate of a function $f(x)$ is defined as
\[ f^*(y) = \sup_x \left( \langle x, y \rangle - f(x) \right). \]
Here $\langle x, y \rangle = x^T y$, the standard inner product on $\mathbb{R}^n$. When $f(x)$ is differentiable and smooth, the conjugate function $f^*(y)$ achieves the maximum value at $x^* = \nabla f(x)$. This is a special instance of the ``Legendre" transform \cite{boyd}. To achieve ``safety" guarantees in reinforcement learning, such as ensuring a robot learning a task never moves into dangerous values of the parameter space, we need to ensure that when domain constraints are not violated. We use Bregman divergences \cite{bregman} to ensure that safety constraints are adhered to, where the projection is defined as:
\[ \Pi^\Phi_{\cal X}(y) = \mbox{argmin}_{{\cal X} \cap {\cal D}} D_\Phi(x,y). \]
A distance generating function $\Phi(x)$ is defined as a strongly convex function which is differentiable. Given such a function $\Phi$, the Bregman divergence associated with it is defined as:
\begin{equation}
\label{bregmand1}
D_\Phi(x,y) = \Phi(x) - \Phi(y) - \langle \nabla \Phi(y), x - y \rangle
\end{equation}
Intuitively, the Bregman divergence measures the difference between the value of a strongly convex function $\Phi(x)$ and the estimate derived from the first-order Taylor series expansion at $\Phi(y)$. Many widely used distance measures turn out to be special cases of Bregman divergences, such as Euclidean distance (where $\Phi(x) = \frac{1}{2} \| x\|^2$ ) and Kullback Liebler divergence (where $\Phi(x) = \sum_i x_i \log_2 x_i$, the negative entropy function). In general, Bregman divergences are non-symmetric, but projections onto a convex set with respect to a Bregman divergence is well-defined.
\subsection{Mirror Descent, Extragradient, and Mirror Prox Methods}
The framework of {\em mirror descent} \cite{nemirovski-yudin:book,Beck:2003p2359} plays a central role in our framework, which includes not just the original mirror descent method, but also the mirror-prox method \cite{nemirovski2005prox}, which generalizes the {\em extragradient} method to non-Euclidean geometries \cite{extragradient:1976}. Figure~\ref{mdfig} illustrates the mirror descent method, and Figure~\ref{egfig} illustrates the extragradient method.
\begin{figure}[ht]
\centering
\begin{minipage}[t]{0.7\textwidth}
\includegraphics[height=0.35\textheight,width=4in]{mdfig.pdf}
\end{minipage}
\caption{The mirror descent method. This figure is adapted from \cite{bubeck}.}
\label{mdfig}
\end{figure}
The extragradient method was developed to solve {\em variational inequalities} (VIs), a beautiful generalization of optimization. Variational inequalities, in the infinite-dimensional setting, were originally proposed by Hartman and Stampacchia \cite{hartman-stampacchia:acta} in the mid-1960s in the context of solving partial differential equations in mechanics. Finite-dimensional VIs rose in popularity in the 1980s partly as a result of work by Dafermos \cite{dafermos}, who showed that the traffic network equilibrium problem could be formulated as a finite-dimensional VI. This advance inspired much follow-on research, showing that a variety of equilibrium problems in economics, game theory, sequential decision-making etc. could also be formulated as finite-dimensional VIs -- the books by Nagurney \cite{nagurney:vibook} and Facchinei and Pang \cite{facchinei-pang:vi} provide a detailed introduction to the theory and applications of finite-dimensional VIs. While we leave the full explication of the VI approach to reinforcement learning to a subsequent monograph, we discuss in the last chapter a few intriguing aspects of this framework that is now the subject of another investigation by our group. A VI(F,K) is specified by a vector field F and a feasible set K. Solving a VI means finding an element $x^*$ within the feasible set K where the vector field $F(x^*)$ is pointed inwards and makes an acute angle with all vectors $x - x^*$. Equivalently, $-F(x^*)$ belongs in the normal cone of the convex feasible set K at the point $x^*$. Any optimization problem reduces to a VI, but the converse is only true for vector fields F whose Jacobians are symmetric. A more detailed discussion of VIs is beyond the scope of this paper, but a longer summary is given in Chapter~\ref{vis}.
In Figure~\ref{egfig}, the concept of extragradient is illustrated. A simple way to understand the figure is to imagine the vector field F here is defined as the gradient $\nabla f(x)$ of some function being minimized. In that case, the mapping $-F(x_k)$ points as usual in the direction of the negative gradient. However, the clever feature of extragradient is that it moves not in the direction of the negative gradient at $x_k$, but rather in the direction of the negative gradient at the point $y_k$, which is the projection of the original gradient step onto the feasible set K. We will see later how this property of extragradient makes its appearance in accelerating gradient temporal difference learning algorithms, such as TDC \cite{Sutton09fastgradient-descent}.
\begin{figure}[ht]
\centering
\begin{minipage}[t]{0.7\textwidth}
\includegraphics[height=0.25\textheight,width=4in]{eg-one-step.pdf}
\end{minipage}
\caption{The extragradient method.}
\label{egfig}
\end{figure}
The mirror-prox method generalizes the extragradient method to non-Euclidean geometries, analogous to the way mirror descent generalizes the regular gradient method. The mirror-prox algorithm (MP) \cite{nemirovski2005prox} is a first-order approach that is able to solve saddle-point problems at a convergence rate of $O(1/t)$. The MP method plays a key role in our framework as our approach extensively uses the saddle point reformulation of reinforcement learning developed by us \cite{ROTD:NIPS2012}. Figure~\ref{mirrorprox} illustrates the mirror-prox method.
\begin{figure}[ht]
\centering
\begin{minipage}[t]{0.7\textwidth}
\includegraphics[height=0.35\textheight,width=4in]{mirrorprox.pdf}
\end{minipage}
\caption{The mirror prox method. This figure is adapted from \cite{bubeck}.}
\label{mirrorprox}
\end{figure}
\subsection{Proximal Operators}
We now review the concept of proximal mappings, and then describe its relation to the mirror descent framework. The proximal mapping associated with a convex function $h$ is defined as:
\[ \mbox{prox}_h(x) = \mbox{argmin}_{u \in X} \left( h(u) +\frac{1}{2} \| u - x \|^2 \right) \]
If $h(x) = 0$, then $\mbox{prox}_h(x) = x$, the identity function. If $h(x) = I_C(x)$, the indicator function for a convex set $C$, then $\mbox{prox}_{I_C}(x) = \Pi_C(x)$, the projector onto set $C$. For learning sparse representations, the case when $h(w) = \lambda \| w\|_1$ (the $L_1$ norm of $w$) is particularly important. In this case:
\begin{equation}
\label{l1prox}
\mbox{prox}_h(w)_i = \begin{cases} w_i - \lambda, & \mbox{if } w_i > \lambda \\ 0, & \mbox{if } |w_i| \leq \lambda \\ w_i + \lambda, & \mbox{otherwise} \end{cases}
\end{equation}
%
An interesting observation follows from noting that the projected subgradient method can be written equivalently using the proximal mapping as:
\[ w_{k+1} = \mbox{argmin}_{w \in X} \left( \langle w, \partial f(w_k) \rangle + \frac{1}{2 \alpha_k} \| w - w_k \|^2 \right) \]
where $X$ is a closed convex set. An intuitive way to understand this equation is to view the first term as requiring the next iterate $w_{k+1}$ to move in the direction of the (sub) gradient of $f$ at $w_k$, whereas the second term requires that the next iterate $w_{k+1}$ not move too far away from the current iterate $w_k$.
With this introduction, we can now introduce the main concept of {\em mirror descent}, which was originally proposed by Nemirovksi and Yudin \cite{nemirovski-yudin:book}. We follow the treatment in \cite{Beck:2003p2359} in presenting the mirror descent algorithm as a nonlinear proximal method based on a distance generator function that is a Bregman divergence \cite{bregman}. The general mirror descent procedure can thus be defined as:
\begin{equation}
\label{md}
w_{k+1} = \mbox{argmin}_{w \in X} \left( \langle w, \partial f(w_k) \rangle + \frac{1}{\alpha_k} D_\psi(w,w_k) \right)
\end{equation}
%
The solution to this optimization problem can be stated succinctly as the following generalized gradient descent algorithm, which forms the core procedure in mirror descent:
%
\begin{equation}
\label{md-update}
w_{k+1} = \nabla \psi^* \left( \nabla \psi(w_k) - \alpha_k \partial f(w_k) \right)
\end{equation}
%
An intuitive way to understand the mirror descent procedure specified in Equation~\ref{md-update} is to view the gradient update in two stages: in the first step, the gradient is computed in the dual space using a set of auxiliary weights $\theta$, and subsequently the updated auxilary weights are mapped back into the primal space $w$. Mirror descent is a powerful first-order optimization method that is in some cases ``optimal" in that it leads to low regret. One of the earliest and most successful applications of mirror descent is Positron Emission Tomography (PET) imaging, which involves minimizing a convex function over the unit simplex $X$. It is shown in \cite{BenTal:2001p2356} that the mirror descent procedure specified in Equation~\ref{md-update} with the Bregman divergence defined by the {\em p-norm} function \cite{gentile:mlj} can outperform regular projected subgradient method by a factor $\frac{n}{\log n}$ where $n$ is the dimensionality of the space. For high-dimensional spaces, this ratio can be quite large. We will discuss below specific choices of Bregman divergences in the target application of this framework to reinforcement learning.
\subsection{Operator Splitting Strategies}
In our framework, a key insight used to derive a true stochastic gradient method for reinforcement learning is based on the powerful concept of {\em operator splitting} \cite{operator-splitting,douglas-rachford}. Figure~\ref{operator-splitting} illustrates this concept for the {\em convex feasibility} problem, where we are given a collection of convex sets, and have to find a point in their intersection. This problem originally motivated the development of Bregman divergences \cite{bregman}. The convex feasibility problem is an example of many real-world problems, such as 3D voxel reconstruction in brain imaging \cite{BenTal:2001p2356}, a high-dimensional problem that mirror descent was originally developed for. To find an element in the common intersection of two sets $A$ and $B$ in Figure~\ref{operator-splitting}, a standard method called {\em alternating projections} works as follows. Given an initial point $x_0$, the first step projects it to one of the two convex sets, say $A$, giving the point $\Pi_A(x_0)$. Since $A$ is convex, this is a uniquely defined point. The next step is to project the new point on the second set $B$, giving the next point $\Pi_B(\Pi_A(x_0))$. The process continues, ultimately leading to the desired point common to the two sets. Operator splitting studies a generalized version of this problem, where the projection problem is replaced by the proximal operator problem, as described above. Many different operator splitting strategies have been developed, such as Douglas Rachford splitting \cite{douglas-rachford}, which is a generalization of widely used distributed optimization methods like Alternating Direction Method of Multipliers \cite{boyd:admm}. We will see later that using a sophisticated type of operator splitting strategy, we can address the problem of off-policy temporal difference learning.
\begin{figure}[ht]
\centering
\begin{minipage}[t]{0.7\textwidth}
\includegraphics[height=0.35\textheight,width=4in]{operator-splitting.pdf}
\end{minipage}
\caption{Operator splitting strategy for the convex feasibility problem.}
\label{operator-splitting}
\end{figure}
\section{Illustrating the Solution}
Now that we have described the broad elements of our framework, we give a few select examples of the tangible solutions that emerge to the problem of designing safe, reliable, and stable reinforcement learning algorithms. We pick three cases: how to design a ``safe" reinforcement learning method; how to design a ``true" stochastic gradient reinforcement learning method; and finally, how to design a ``robust" reinforcement learning method that does not overfit its training experience.
\section{Safe Reinforcement Learning}
\begin{figure}
\centering
\includegraphics[width=.5\columnwidth]{uBot.jpg}
\caption{The uBot-5 is a 11 degree of freedom mobile manipulator developed at the Laboratory of Perceptual Robotics (LPR) at the University of Massachusetts, Amherst \cite{Deegan2010,Kuindersma2009}. How can we design a ``safe" reinforcement learning algorithm which is guaranteed to ensure that policy learning will not violate pre-defined constraints such that such robots will operate in dangerous regions of the control parameter space? Our framework provides a key solution, based on showing an equivalence between mirror descent and a previously well-studied but unrelated algorithm called {\em natural gradient} \cite{Amari1998b}.}
\label{ch1fig:uBot}
\end{figure}
Figure~\ref{ch1fig:uBot} shows a complex high-degree of freedom humanoid robot. Teaching robots complex skills is a challenging problem, particularly since reinforcement learning not only may take a long time, but also because it may cause such robots to operate in dangerous regions of the parameter space. Our proposed framework solves this problem by establishing a key technical result, stated below, between mirror descent and the well-known, but previously unrelated, class of algorithms called natural gradient \cite{Amari1998b}. We develop the projected natural actor critic (PNAC) algorithm, a policy gradient method that exploits this equivalence to yield a safe method for training complex robots using reinforcement learning. We explain the significance of the below result connecting mirror descent and natural gradient methods later in this paper when we describe a novel class of methods called projected natural actor critic (PNAC).
\begin{thm}
The natural gradient descent update at step $k$ with metric tensor $G_k \triangleq G(x_k)$:
\begin{equation}
x_{k+1} = x_k - \alpha_k G_k^{-1}\nabla f(x_k),
\label{eq:NG2}
\end{equation}
is equivalent to the mirror descent update at step $k$, with $\psi_k(x)=(\sfrac{1}{2})x^\intercal G_k x$.
\end{thm}
\section{True Stochastic Gradient Reinforcement Learning}
First-order temporal difference learning is a widely used class of techniques
in reinforcement learning. Although many more sophisticated methods have been developed over the past three decades,
such as least-squares based
temporal difference approaches, including LSTD \cite{bradtke-barto:LSTD},
LSPE \cite{bertsekas-nedic} and LSPI \cite{lagoudakis:jmlr}, first-order temporal difference
learning algorithms may scale more gracefully to high dimensional problems.
Unfortunately, the initial class of TD methods was known to converge only when samples are drawn
``on-policy". This motivated the development of the gradient TD (GTD) family of methods \cite{FastGradient:2009}. A crucial step in the development of our framework was the development of a novel saddle-point framework for sparse regularized GTD \cite{ROTD:NIPS2012}. However, there have been several unresolved questions
regarding the current off-policy TD algorithms.
(1) The first is the convergence rate of these algorithms. Although
these algorithms are motivated from the gradient of an objective function
such as mean-squared projected Bellman error (MSPBE) and NEU \cite{FastGradient:2009}, they are not true stochastic gradient methods
with respect to these objective functions, as pointed out in \cite{szepesvari2010algorithms}, which make the convergence rate and error bound analysis difficult, although asymptotic analysis
has been carried out using the ODE approach. (2) The second concern is
regarding acceleration. It is believed that TDC performs the best so far of
the GTD family of algorithms. One may intuitively ask if there are any gradient TD
algorithms that can outperform TDC. (3) The third concern is regarding compactness
of the feasible set $\theta$. The GTD family of algorithms all assume
that the feasible set $\theta$ is unbounded, and if the feasible
set $\theta$ is compact, there is no theoretical analysis and convergence
guarantee. (4) The fourth question is on regularization: although
the saddle point framework proposed in \cite{ROTD:NIPS2012} provides an online regularization framework
for the GTD family of algorithms, termed as RO-TD, it is based on the
inverse problem formulation and is thus not quite explicit. One further
question is whether there is a more straightforward algorithm, e.g,
the regularization is directly based on the MSPBE and NEU objective
functions.
Biased sampling is a well-known problem in reinforcement learning.
Biased sampling is caused by the stochasticity of the policy wherein
there are multiple possible successor states from the current state
where the agent is. If it is a deterministic policy, then there will
be no biased sampling problem. Biased sampling is often caused
by the product of the TD errors, or the product of TD error and the
gradient of TD error w.r.t the model parameter $\theta$. There are
two ways to avoid the biased sampling problem, which can be categorized
into double sampling methods and two-time-scale stochastic approximation
methods.
In this paper, we propose a novel approach to TD algorithm design in reinforcement learning, based on
introducing the {\em proximal splitting} framework \cite{PROXSPLITTING2011}. We show that the GTD family of algorithms are true stochastic gradient descent (SGD) methods, thus making their convergence
rate analysis available. New accelerated off-policy algorithms are
proposed and their comparative study with RO-TD is carried out to show
the effectiveness of the proposed algorithms. We also show that primal-dual
splitting is a unified first-order optimization framework to solve the
biased sampling problem. Figure~\ref{ch1fig:star} compares the performance of our newly designed off-policy methods compared to previous methods, like TDC and GTD2 on the
classic 5-state Baird counterexample. Note the significant improvement of TDC-MP over TDC: the latter converges much more slowly, and has much higher variance. This result is validated not only by experiments, but also by a detailed theoretical analysis of sample convergence, which goes beyond the previous asymptotic convergence analysis of off-policy methods.
\begin{figure}
\centering
\begin{minipage}{1.15\textwidth}
\includegraphics[width= .45\textwidth, height=2.25in]{baird_1.pdf}
\includegraphics[width= .45\textwidth, height=2.25in]{baird_tdc.pdf}
\end{minipage}
\caption{Off-Policy Convergence Comparison. Our proposed methods, TDC-MP and GTD2-MP, appear to significantly outperform previous methods, like TDC and GTD2 on a simple
benchmark MDP.}
\label{ch1fig:star}
\end{figure}
\section{Sparse Reinforcement Learning using Mirror Descent}
How can we design reinforcement learning algorithms that are robust to overfitting? In this paper we explore a new framework for (on-policy convergent)
TD learning algorithms based on mirror descent and related
algorithms. Mirror descent can be viewed as an enhanced gradient method,
particularly suited to minimization of convex functions in high-dimensional
spaces. Unlike traditional temporal difference learning methods, mirror descent temporal difference learning undertakes
updates of weights in both the dual space and primal space,
which are linked together using a Legendre transform. Mirror descent
can be viewed as a proximal algorithm where the distance-generating
function used is a Bregman divergence. We will present a new class of {\em proximal-gradient}
based temporal-difference (TD) methods based on different
Bregman divergences, which are more powerful than regular TD learning.
Examples of Bregman divergences that are studied include $p$-norm
functions, and Mahalanobis distance based on the covariance of sample
gradients. A new family of sparse mirror-descent reinforcement learning
methods are proposed, which are able to find sparse fixed-point of
an $l_{1}$-regularized Bellman equation at significantly less computational
cost than previous methods based on second-order matrix methods. Figure~\ref{ch1mcar-experiment} illustrates a sample result, showing how the mirror descent variant of
temporal difference learning results in faster convergence, and much lower variance (not shown) on the classic mountain car task \cite{sutton-barto:book}.
\begin{figure}[tbh]
\centering
\begin{minipage}[t]{0.8\textwidth}
\includegraphics[height=2.5in, width=4in]{mcar-mdtd-vs-regular-td-16-fourier.pdf}
\end{minipage}
\caption{Comparing mirror-descent TD using the p-norm link function with $16$ tunable Fourier bases with regular TD for the mountain car task. }
\label{ch1mcar-experiment}
\end{figure}
\section{Summary}
We provided a brief overview of our proposed primal-dual framework for reinforcement learning. The fundamentally new idea underlying the approach is the systematic use of mirror maps to
carry out temporal difference updates, not in the original primal space, but rather in a {\em dual} space. This technical device, as we will show in subsequent chapters, provides for a number
of significant advantages. By choosing the mirror map carefully, we can generalize popular methods like natural gradient based actor-critic methods, and provide safety guarantees. We
can design more robust temporal difference learning methods that are less prone to overfitting the experience of an agent. Finally, we can exploit proximal mappings to design a rich variety
of true stochastic gradient methods. These advantages, when combined, provide a compelling case for the fundamental correctness of our approach. However, much remains to be done in more fully
validating the proposed framework on large complex real-world applications, as well as doing a deeper theoretical analysis of our proposed approach. These extensions will be the subject of ongoing
research by us in the years ahead.
\chapter{Background}\label{chapter:Background}
In this chapter we introduce relevant background material that form the two cornerstones of this paper: reinforcement learning and first-order stochastic composite optimization. The Markov decision process (MDP) model, value function approximation and some basics of reinforcement learning are also introduced. For stochastic composite optimization, we first introduce the problem formulation, and then introduce some tools such as proximal gradient method, mirror descent, etc.
\section{Reinforcement Learning}
\subsection{MDP}
The learning environment for decision-making is generally modeled
by the well-known \textbf{\emph{Markov Decision Process}}\cite{puterman} $M=(S,A,P,R,\gamma)$,
which is derived from a Markov chain.
\begin{defn}
(Markov Chain)\label{def:mc}: A \emph{Markov Chain} is a stochastic process defined as $M=(S,P)$. At each time step $t=1,2,3,\cdots$,
the agent is in a state ${s_{t}}\in S$, and the state transition
probability is given by the state transition kernel $P:S\times S\to\mathbb{{R}}$
satisfying $||P|{|_{\infty}}=1$, where $P({s_{t}}|{s_{t-1}})$ is
the state-transition probability from state $s_{t-1}$ at time step
$t-1$ to the state $s_{t}$ at time step $s_{t}$.
\end{defn}
A Markov decision process (MDPs) is comprised of a set of states $S$,
a set of (possibly state-dependent) actions $A$ ($A_{s}$), a dynamical
system model comprised of the transition probabilities $P_{ss'}^{a}$
specifying the probability of transition to state $s'$ from state
$s$ under action $a$, and a reward model $R$.
\begin{defn}
(Markov Decision Process)\label{def:mdp}\cite{puterman}: A Markov
Decision Process is a tuple $(S,A,P,R,\gamma)$ where $S$ is a finite
set of states, $A$ is a finite set of actions, $P:S\times A\times S\to[0,1]$
is the transition kernel, where $P(s,a,s')$ is the probability of
transmission from state $s$ to state $s'$ given action $a$, and
reward $r:S\times A\to{\mathbb{R}{}^{+}}$ is a reward function, $0\leq\gamma<1$
is a discount factor.
\end{defn}
\subsection{Basics of Reinforcement Learning}
A policy $\pi:S\rightarrow A$ is a deterministic (stochastic) mapping
from states to actions.
\begin{defn}
\noindent\label{def:pol}(Policy): A deterministic stationary policy
$\pi:S\to A$ assigns an action to each state of the Markov decision
process. A stochastic policy $\pi:S\times A\to[0,1]$.
\end{defn}
Value functions are used to compare and evaluate the performance of
policies.
\begin{defn}
\noindent\label{def:value}(Value Function): A value function w.r.t
a policy $\pi$ termed as ${V^{\pi}}:S\to\mathbb{R}$ assigns each state the
\emph{expected sum of discounted} rewards
\end{defn}
\begin{equation}
V^{\pi}=\mathbb{E}\left[{\sum\limits _{i=1}^{t}{{\gamma^{i-1}}{r_{i}}}}\right]
\end{equation}
The goal of reinforcement learning is to find a (near-optimal) policy
that maximizes the value function. $V^{\pi}$ is a fixed-point of
the Bellman equation
\begin{equation}
{V^{\pi}}({s_{t}})=\mathbb{E}\left[{r({s_{t}},{\pi(s_{t})})+\gamma{V^{\pi}}({s_{t+1}})}\right]\label{eq:bellman_eq}
\end{equation}
Equation (\ref{eq:bellman_eq}) can be written in a concise form by
introducing the Bellman operator $T^{\pi}$ w.r.t a policy $\pi$
and denoting the reward vector as $R^{\pi}\in{\mathbb{R}^{n}}$ where ${R^{\pi}_{i}}=\mathbb{E}[r({s_{i}},\pi(s_{i}))]$.
\begin{equation}
V^{\pi}=T^{\pi}(V^{\pi})=R^{\pi}+\gamma P^{\pi}V^{\pi}\label{eq:bellman_op}
\end{equation}
Any optimal policy $\pi^{*}$ defines the unique optimal value function
$V^{*}$ that satisfies the nonlinear system of equations:
\begin{equation}
V^{^{*}}(s)=\max_{a}\sum_{s'}P_{ss'}^{a}\left(R_{ss'}^{a}+\gamma V^{*}(s')\right)\label{eq:Vopt}
\end{equation}
\subsection{Value Function Approximation}
The most popular and widely used RL method is temporal difference
(TD) learning \cite{sutton:td}. TD learning is a stochastic approximation
approach to solving Equation (\ref{eq:Vopt}). The {\em state-action
value} $Q^{*}(s,a)$ represents a convenient reformulation of the
value function, defined as the long-term value of performing $a$
first, and then acting optimally according to $V^{*}$:
\begin{equation}
Q^{*}(s,a)=\mathbb{E}\left(r_{t+1}+\gamma\max_{a'}Q^{*}(s_{t+1},a')|s_{t}=s,a_{t}=a\right)\label{qvalues}
\end{equation}
where $r_{t+1}$ is the actual reward received at the next time step,
and $s_{t+1}$ is the state resulting from executing action $a$ in
state $s_{t}$. The (optimal) action value formulation is convenient
because it can be approximately solved by a temporal-difference (TD)
learning technique called Q-learning \cite{watkins:phd}. The simplest
TD method, called TD($0$), estimates the value function associated
with the fixed policy using a normal stochastic gradient iteration,
where $\delta_{t}$ is called temporal difference error:
\begin{equation}
\begin{array}{l}
{V_{t+1}}({s_{t}})={V_{t}}({s_{t}})+{\alpha_{t}}{\delta_{t}}\\
{\delta_{t}}={r_{t}}+\gamma{V_{t}}({s_{t+1}})-{V_{t}}({s_{t}})
\end{array}
\end{equation}
TD($0$) converges to the optimal value function $V^{\pi}$ for policy
$\pi$ as long as the samples are ``on-policy'', namely following
the stochastic Markov chain associated with the policy; and the learning
rate $\alpha_{t}$ is decayed according to the Robbins-Monro conditions
in stochastic approximation theory: $\sum_{t}\alpha_{t}=\infty,\sum_{t}\alpha_{t}^{2}<\infty$
\cite{ndp:book}. When the set of states $S$ is large, it is often
necessary to approximate the value function $V$ using a set of handcrafted
basis functions (e.g., polynomials, radial basis functions, wavelets
etc.) or automatically generated basis functions \cite{mahadevan:mlfnt}.
In linear value function approximation, the value function is assumed
to lie in the linear spanning space of the basis function matrix $\Phi$
of dimension $|S|\times d$, where it is assumed that $d\ll|S|$.
Hence,
\begin{equation}
V^{\pi}\approx V_{\theta}=\Phi\theta
\end{equation}
The equivalent TD($0$) algorithm for linear function approximated
value functions is given as:
\begin{equation}
\begin{array}{l}
{\theta _{t + 1}} = {\theta _t} + {\alpha _t}{\delta _t}\phi ({s_t})\\
{\delta _t} = {r_t} + \gamma \phi {({s_{t + 1}})^T}{\theta _t} - \phi {({s_t})^T}{\theta _t}
\end{array}
\label{td0}
\end{equation}
\section{Stochastic Composite Optimization}
\subsection{Stochastic Composite Optimization Formulation}
Stochastic optimization explores the use of first-order gradient methods
for solving convex optimization problems.
We first give some definitions before moving on to introduce
stochastic composite optimization.
\begin{defn}
\noindent (Lipschitz-continuous Gradient)\label{def:lipschitz}: The
gradient of a closed convex function $f(x)$ is $L$-Lipschitz continuous
if $\exists{L},||\nabla f(x)-\nabla f(y)||\le{L}||x-y||,\forall x,y\in X$.
\end{defn}
\begin{defn}
\noindent (Strong Convexity): A convex function is $\mu-$strongly
convex if $\exists{\mu}$, $\frac{\mu}{2}||x-y|{|^{2}}\le f(y)-f(x)-\left\langle {\nabla f(x),y-x}\right\rangle ,\forall x,y\in X$.
\end{defn}
\noindent \textbf{Remark}: If $f(x)$ is both with $L$-Lipschitz
continuous gradient and $\mu$-strongly convex, then we have $\forall x,y\in X$,
\[
\frac{\mu}{2}||x-y|{|^{2}}\le f(y)-f(x)-\left\langle {\nabla f(x),y-x}\right\rangle \le\frac{L}{2}||x-y|{|^{2}}
\]
\begin{defn}
(Stochastic Subgradient) \label{def:sg}: The stochastic subgradient
for closed convex function $f(x)$ at $x$ is defined as $g(x,\xi_{t})$
satisfying ${\mathbb{E}}[g(x,{\xi_{t}})]=\nabla f(x)\in\partial f(x)$.
Further, we assume that the variance is bounded $\exists\sigma>0$
such that
\begin{equation}
\forall x\in X,{\mathbb{E}}[||g(x,{\xi_{t}})-\nabla f(x)||_{*}^{2}]\le{\sigma^{2}}\label{eq:sgvariance}
\end{equation}
\end{defn}
Here we define the problem of Stochastic Composite Optimization (SCO)\cite{lan2012optimal}:
\begin{defn}
(Stochastic Composite Optimization): A stochastic composite optimization
problem $\mathcal{F}(L,M,\mu,\sigma):\Psi(x)$ on a closed convex
set $X$ is defined as
\begin{equation}
{\min_{x\in X}}\Psi(x)\mathop=\limits ^{def}f(x)+h(x)
\label{sumfg}
\end{equation}
$f(x)$ is a convex function with $L$-Lipschitz continuous gradient and $h(x)$ is a convex Lipschitz continuous function such that
\begin{equation}
|h(x) - h(y)| \le M||x - y||,\forall x,y \in X
\end{equation}
$g(x,\xi_{t})$ is the stochastic subgradient of $\Psi(x)$ defined above with variance bound $\sigma$.
Such $\Psi(x)$ is termed as a $\mathcal{F}(L,M,\mu,\sigma)$ problem.
\end{defn}
\subsection{Proximal Gradient Method and Mirror Descent}
Before we move on to introduce mirror descent, we first introduce
some definitions and notations.
\begin{defn}
\noindent (Distance-generating Function)\cite{MID:2003}: A distance-generating function
$\psi(x)$ is defined as a continuously differentiable $\mu$-strongly
convex function.
$\psi^{*}$ is the Legendre transform of $\psi$,
which is defined as ${\psi^{*}}(y)=\mathop{\sup}\limits _{x\in X}\left({\left\langle {x,y}\right\rangle -\psi(x)}\right)$.
\end{defn}
\begin{defn}
\noindent (Bregman Divergence)\cite{MID:2003}: Given distance-generating function
$\psi$, the Bregman divergence induced by $\psi$ is defined as:
\begin{equation}
D_{\psi}(x,y)=\psi(x)-\psi(y)-\langle\nabla\psi(y),x-y\rangle
\label{bregmand}
\end{equation}
Legendre transform and Bregman divergence have the following properties
\begin{itemize}
\item $\nabla{\psi^{*}}={(\nabla\psi)^{-1}}$
\item ${D_{\psi}}(u,v)={D_{{\psi^{*}}}}(\nabla\psi(u),\nabla\psi(v))$
\item $\nabla{D_{\psi}}(u,v)=\nabla\psi(u)-\nabla\psi(v)$
\end{itemize}
An interesting choice of the link function $\psi(\cdot)$ is the $(q-1)$-strongly convex function $\psi(\theta)=\frac{1}{2}\|\theta\|_{q}^{2}$,
and ${\psi^{*}}(\tilde{\theta})=\frac{1}{2}||\tilde{\theta}||_{p}^{2}$.
Here, $\|\theta\|_{q}=\left(\sum_{j}|\theta_{j}|^{q}\right)^{\frac{1}{q}}$,
and $p$ and $q$ are conjugate numbers such that $\frac{1}{p}+\frac{1}{q}=1$
\cite{gentile2003robustness}. $\theta$ and $\tilde{\theta}$ are
conjugate variables in primal space and dual space, respectively .
\begin{eqnarray}
\mathop{\nabla\psi}\limits _{\theta\to\tilde{\theta}}{(\theta)_{j}} & = & \frac{{{\rm {sign}}({\theta_{j}})|{\theta_{j}}{|^{q-1}}}}{{||\theta||_{q}^{q-2}}}\nonumber \\
\;{\mathop{\;\nabla\psi}\limits _{\tilde{\theta}\to\theta}}^{*}{(\tilde{\theta})_{j}} & = & \frac{{{\rm {sign}}({{\tilde{\theta}}_{j}})|{{\tilde{\theta}}_{j}}{|^{p-1}}}}{{||\tilde{\theta}||_{p}^{p-2}}}
\end{eqnarray}
Also it is worth noting that when $p=q=2$, the Legendre transform
is the identity mapping.
\end{defn}
We now introduce the concept of \emph{proximal mapping}, and then
describe the mirror descent framework. The proximal mapping associated
with a convex function $h(x)$ is defined as:
\begin{equation}
prox_{h}(x)=\arg\mathop{\min}\limits _{u\in X}(h(u)+\frac{1}{2}{\left\Vert {u-x}\right\Vert ^{2}})\label{eq:0}
\end{equation}
In the case of $h(x)=\rho{\left\Vert x\right\Vert _{1}}(\rho>0)$,
which is particularly important for sparse feature selection, the
proximal operator turns out to be the soft-thresholding operator ${S_{\rho}}(\cdot)$,
which is an \emph{entry-wise} shrinkage operator that moves a point
towards zero, i.e.,
\begin{equation}
pro{x_{h}}{(x)_{i}}={S_{\rho}}{(x)_{i}}={\rm {sign}}({x_{i}})\max({|x_{i}-\rho|},0)\label{softthresholding}
\end{equation}
where $i$ is the index, and $\rho$ is a threshold. With this background,
we now introduce the proximal gradient method. At each iteration, the
optimization sub-problem of Equation (\ref{sumfg}) can be rewritten
as
\begin{equation}
{x_{t+1}}=\arg\mathop{\min}\limits _{u\in X}(h(u)+\langle\nabla{f_{t}},u\rangle+\frac{1}{{2{\alpha_{t}}}}{\left\Vert {u-x_{t}}\right\Vert ^{2}})\label{eq:proxgradient}
\end{equation}
If computing $prox_{h}$ is not expensive, then computation of Equation
(\ref{sumfg}) is of the following formulation, which is called the \emph{proximal
gradient method}
\begin{equation}
{x_{t+1}}=pro{x_{{\alpha_{t}}h}}\left({{x_{t}}-{\alpha_{t}}\nabla f({x_{t}})}\right)\label{eq:linearproximal}
\end{equation}
where ${\alpha_{t}}>0$ is stepsize, constant or determined by line
search. The mirror descent \cite{MID:2003} algorithm is a generalization
of classic gradient descent, which has led to developments of new
more powerful machine learning methods for classification and regression.
Mirror descent can be viewed as an enhanced gradient method, particularly
suited to minimization of convex functions in high-dimensional spaces.
Unlike traditional gradient methods, mirror descent undertakes gradient
updates of weights in the dual space, which is linked together with
the primal space using a Legendre transform. Mirror descent can be
viewed as a proximal algorithm where the distance-generating function
used is a Bregman divergence w.r.t the distance-generating function
$\psi$, and thus the optimization problem is
\begin{equation}
prox_{h}(x)=\arg\mathop{\min}\limits _{u\in X}(h(u)+D_{\psi}(u,x))\label{eq:0breg}
\end{equation}
The solution to this optimization problem of Equation (\ref{eq:0breg})
forms the core procedure of \emph{mirror descent} as a generalization
of Equation (\ref{eq:proxgradient})
\begin{equation}
{x_{t+1}}=\arg\mathop{\min}\limits _{u\in X}(h(u)+\langle\nabla{f_{t}},u\rangle+\frac{1}{{\alpha_{t}}}{D_{\psi}}(u,{x_{t}}))\label{eq:proxgradientmid}
\end{equation}
which is a nonlinear extension of Equation(\ref{eq:linearproximal})
\begin{equation}
{x_{t+1}}=\nabla{\psi^{*}}\left({pro{x_{{\alpha_{t}}h}}\left({\nabla\psi({x_{t}})-{\alpha_{t}}\nabla f({x_{t}})}\right)}\right)\label{eq:MIDSA}
\end{equation}
Mirror descent has become the cornerstone of many online $l_{1}$
regularization approaches such as in \cite{ShalevShwartz:jmlr}, \cite{RDA:2009}
and \cite{COMID:2010}.
\subsection{Dual Averaging}
Regularized dual averaging (RDA) \cite{RDA:2009} is a variant of
Dual averaging (DA) with ``simple'' regularizers, such as $l_{1}$
regularization. DA method is strongly related to cutting-plane methods.
Cutting-plane methods formulate a polyhedral lower bound model of
the objective function where each gradient from past iterations contributes
a supporting hyperplane w.r.t its corresponding previous iteration,
which is often expensive to compute. The DA method approximates this
lower bound model with an approximate (possibly not supporting) lower
bound hyperplane with the averaging of all the past gradients \cite{gaojianfeng:2013}.
We now explain RDA from the proximal gradient perspective. Thus far,
the proximal gradient methods we have described in Equation (\ref{eq:MIDSA})
adjust the weights to lie in the direction of the current gradient
$\nabla f_{t}$. Regularized dual averaging methods (RDA) uses a (weighted)
averaging of gradients, which explain their name. Compared with Equation
(\ref{eq:MIDSA}), the main difference is the average (sub)gradient
$\nabla{{\bar{f}}_{t}}$ is used, where $\nabla{\bar{f}_{t}}=\frac{1}{t}\sum\limits _{i=1}^{t}{\nabla{f_{i}}}$.
The equivalent space-efficient recursive representation is
\begin{equation}
\nabla{{\bar{f}}_{t}}=\frac{{t-1}}{t}\nabla{{\bar{f}}_{t-1}}+\frac{1}{t}\nabla{f_{t}}\label{barft-1}
\end{equation}
The generalized mirror-descent proximal gradient formulation of RDA
iteratively solves the following optimization problem at each step:
\begin{equation}
{x_{t+1}}=\arg\mathop{\min}\limits _{x\in X}\left\{ {\left\langle {x,\nabla{{\bar{f}}_{t}}}\right\rangle +h(x)+\frac{1}{{\alpha_{t}}}{D_{\psi}}(x)}\right\} \label{eq:RDAproximal-1}
\end{equation}
Note that different from Equation (\ref{eq:proxgradientmid}), besides
the averaging gradient ${\nabla{{\bar{f}}_{t}}}$ is used instead
of ${\nabla{{f}_{t}}}$, a global origin-centered stabilizer ${{D_{\psi}}(x)}$
is used. RDA with local stabilizer can be seen in \cite{mcmahan2011follow}.
There are several advantages of RDA over other competing methods in
regression and classification problems. The first is the sparsity
of solution when the penalty term is $h(x)=\rho||x||_{1}$. Compared
with other first-order $l_{1}$ regularization algorithms of the mirror-descent
type, including truncated gradient method \cite{Lanford:2008} and
SMIDAS \cite{SMIDAS:2009}, RDA tends to produce sparser solutions
in that the RDA method is more aggressive on sparsity than many other
competing approaches. Moreover, many optimization problems can be
formulated as composite optimization, e.g., a smooth objective component
in conjunction with a global non-smooth regularization function. It
is worth noting that problems with non-smooth regularization functions
often lead to solutions that lie on a low-dimensional supporting data
manifold, and regularized dual averaging is capable of identifying
this manifold, and thus bringing the potential benefit of accelerating
convergence rate by searching on the low-dimensional manifold after
it is identified, as suggested in \cite{lee2012manifold}. Moreover, the finite iteration behavior
of RDA is much better than SGD in practice.
\subsection{Extragradient}
The extragradient method was first proposed by Korpelevich\cite{extragradient:1976} as a relaxation
of ordinary gradient descent to solve variational
inequality (VI) problems. Conventional ordinary gradient descent can
be used to solve VI problems only if some strict restrictions
such as strong monotonicity of the operator or compactness of the
feasible set are satisfied. The extragradient method was proposed to solve
VIs to relax the aforementioned strict restrictions. The
essence of extragradient methods is that instead of moving along the
steepest gradient descent direction w.r.t the initial point in each iteration, two steps,
i.e., a extrapolation step and a gradient descent step, are taken. In
the extrapolation step, a step is made along the steepest gradient descent
direction of the initial point, resulting in an intermediate point
which is used to compute the gradient. Then the gradient descent step
is made \emph{from} the initial point in the direction of the gradient
w.r.t the intermediate point. The extragradient take steps as follows
\begin{equation}
\begin{array}{l}
{x_{t+\frac{1}{2}}}={\Pi_{X}}\left({{x_{t}}-\alpha_{t}\nabla f({x_{t}})}\right)\\
{x_{t+1}}={\Pi_{X}}\left({{x_{t}}-\alpha_{t}\nabla f({x_{t+\frac{1}{2}}})}\right)
\end{array}\label{eq:eg}
\end{equation}
$\Pi_{X}(x)=\mbox{argmin}_{y\in X}\|x-y\|^{2}$ is the
projection onto the convex set $X$, and $\alpha_{t}$ is a stepsize.
Convergence of the iterations of Equation (\ref{eq:eg}) is guaranteed
under the constraints $0<{\alpha_{t}}<\frac{1}{{\sqrt{2}L}}$\cite{nemirovski2005prox}, where
$L$ is the Lipschitz constant for $\nabla f({x})$.
\subsection{Accelerated Gradient}
Nesterov's seminal work on accelerated gradient (AC) enables deterministic
smooth convex optimization to reach its optimal convergence rate $O(\frac{L}{{N^{2}}})$.
The AC method consists of three major steps: an interpolation step,
a proximal gradient step and a weighted averaging step. During each
iteration,
\begin{eqnarray}
{y_{t}} & = & {\alpha_{t}}{x_{t-1}}+(1-{\alpha_{t}}){z_{t-1}}\nonumber \\
{x_{t}} & = & \arg\mathop{\min}\limits _{x}\left\{ {\left\langle {x,\nabla f({y_{t}})}\right\rangle +h(x)+\frac{1}{{\beta_{t}}}{D_{\psi}}(x,{x_{t-1}})}\right\} \nonumber \\
{z_{t}} & = & {\alpha_{t}}{x_{t}}+(1-{\alpha_{t}}){z_{t-1}}
\end{eqnarray}
It is worth noting that in the proximal gradient step, the stabilizer
makes $x_{t}$ start from $x_{t-1}$, and go along the gradient
descent direction of $\nabla f({y_{t})}$, which is quite similar
to extragradient. The essence of Nesterov's accelerated gradient
method is to carefully select the prox-center for proximal gradient
step, and the selection of two stepsize sequences $\{{\alpha_{t}},{\beta_{t}}\}$
where $\alpha_{t}$ is for interpolation and averaging, $\beta_{t}$
is for proximal gradient. Later work and variants of Nesterov's method
utilizing the strong convexity of the loss function with Bregman divergence
are summarized in \cite{tseng2008accelerated}. Recently, the extension
of accelerated gradient method from deterministic smooth convex optimization
to stochastic composite optimization, termed as AC-SA, is studied
in \cite{lan2012optimal}.
\section{Subdifferentials and Monotone Operators}
We introduce the important concept of a subdifferential.
\begin{definition}
The {\em subdifferential} of a convex function $f$ is defined as the set-valued mapping $\partial f$:
\[ \partial f(x) = \{v \in \mathbb{R}^n : f(z) \geq f(x) + v^T (z - x), \forall z \in \mbox{dom}(f) \]
\end{definition}
A simple example of a subdifferential is the {\em normal cone}, which is the subdifferential of the indicator function $I_K$ of a convex set $K$ (defined as $0$ within the set and $+\infty$ outside). More formally, the normal cone $N_K(x^*)$ at the vector $x^*$ of a convex set $K$ is defined as $N_K(x^*) = \{y \in \mathbb{R}^n | y^T(x - x^*) \leq 0, \forall x \in K \}$. Each vector $v \in \partial f(x)$ is referred to as the {\em subgradient} of $f$ at $x$.
An important property of closed proper convex functions is that their subdifferentials induce a relation on $\mathbb{R}^n$ called a {\em maximal monotone operator} \cite{operator-splitting,rockafellar}.
\begin{definition}
A relation $F$ on $\mathbb{R}^n$ is monotone if
\[ (u - v)^T (x - y) \geq 0 \ \mbox{for all} \ (x, u), (y, v) \in F \]
F is maximal monotone is there is no monotone operator that properly contains it.
\end{definition}
The subdifferential $\partial f$ of a convex function $f$ is a canonical example of a maximal monotone operator. A very general way to formulate optimization problems is {\em monotone inclusion}:
\begin{definition}
Given a monotone operator $F$, the monotone inclusion problem is to find a vector $x$ such that $0 \in F(x)$. For example, given a (subdifferentiable) convex function $f$, finding a vector $x^*$ that minimizes $f$ is equivalent to solving the monotone inclusion problem $0 \in \partial f(x^*)$.
\end{definition}
\section{Convex-concave Saddle-Point First Order Algorithms}
\label{ch2sec:saddle}
A key novel contribution of our paper is a convex-concave saddle-point formulation for reinforcement learning.
A convex-concave saddle-point problem is formulated as follows. Let
$x\in X,y\in Y$, where $X,Y$ are both nonempty closed convex sets, and $f(x):X\to\mathbb{R}$
be a convex function. If there exists a function $\varphi(\cdot,\cdot)$
such that $f(x)$ can be represented as $f(x): = {\sup _{y \in Y}}\varphi (x,y)$,
then the pair $(\varphi,Y)$ is referred as the saddle-point representation
of $f$. The optimization problem
of minimizing $f$ over $X$ is converted into an equivalent convex-concave
saddle-point problem
$SadVal = {\inf _{x \in X}}{\sup _{y \in Y}}\varphi (x,y)$
of $\varphi$ on $X\times Y$. If $f$ is non-smooth
yet convex and well structured, which is not suitable for many existing
optimization approaches requiring smoothness, its saddle-point representation
$\varphi$ is often smooth and convex. The convex-concave saddle-point
problems are, therefore, usually better suited for first-order methods \cite{sra2011optimization}. A comprehensive overview on extending convex minimization to convex-concave saddle-point problems with unified variational inequalities is presented in \cite{ben2005non}.
As an example, consider $f(x) = ||Ax - b|{|_m}$ which admits a bilinear minimax representation
\begin{equation}
f(x): = {\left\| {Ax - b} \right\|_m} = {\max _{{{\left\| y \right\|}_n} < 1}}\left( {\left\langle {y,Ax - b} \right\rangle } \right)
\label{eq:minimax1}
\end{equation}
where $m,n$ are conjugate numbers. Using the approach in \cite{RobustSA:2009}, Equation (\ref{eq:minimax1}) can be solved as
\begin{equation}
{x_{t + 1}} = {x_t} - {\alpha _t}\left\langle {{y_t},A} \right\rangle ,{y_{t + 1}} = {\Pi _{{{\left\| {{y_t}} \right\|}_n} \le 1}}({y_t} + {\alpha _t}(A{x_t} - b))
\label{eq:minimaxFOM}
\end{equation}
where ${\Pi _{{{\left\| {{y_t}} \right\|}_n} \le 1}}$ is the projection operator of $y_t$ onto the unit-$l_n$ ball ${\left\| y \right\|_n} \le {\rm{1}}$,which is defined as
\begin{equation}\label{ch1_l2proj}
{\Pi _{{{\left\| y \right\|}_n} \le 1}}y = \min (1,1/{\left\| y \right\|_n})y,n = 2,{\left( {{\Pi _{{{\left\| y \right\|}_n} \le 1}}y} \right)_i} = \min (1,\frac{1}{{|{y_i}|}}){y_i},n = \infty
\end{equation}
and ${{\Pi _{{{\left\| y \right\|}_\infty} \le 1}}y}$ is an entrywise operator.
\section{Abstraction through Proximal Operators}
A general procedure for solving the monotone inclusion problem, the {\em proximal point algorithm} \cite{rockafellar:prox}, uses the following identities:
\[ 0 \in \partial f(x) \leftrightarrow 0 \in \alpha \partial f(x) \leftrightarrow x \in (I + \alpha \partial (x)) \leftrightarrow x = (I + \alpha \partial f)^{-1}(x) \]
Here, $\alpha > 0$ is any real number. The proximal point algorithm is based on the last fixed point identity, and consists of the following iteration:
\[ x_{k+1} \leftarrow (I + \alpha_k \partial f )^{-1} (x_k) \]
Interestingly, the proximal point method involves the computation of the so-called {\em resolvent} of a relation, defined as follows:
\begin{definition}
The resolvent of a relation F is given as the relation $R_F = (I + \lambda F)^{-1}$, where $\lambda > 0$.
\end{definition}
In the case where the relation $R = \partial f$ of some convex function $f$, the resolvent can be shown to be the {\em proximal mapping} \cite{moreau:prox}, a crucially important abstraction of the concept of projection, a cornerstone of constrained optimization.
\begin{definition}
The proximal mapping of a vector $v$ with respect to a convex function $f$ is defined as the minimization problem:
\[ \mbox{prox}_f(v) = \mbox{argmin}_{x \in K} (f(x) + \| v - x\|^2_2 ) \]
\end{definition}
In the case where $f(x) = I_K(x)$, the indicator function for a convex set $K$, the proximal mapping reduces to the projection $\Pi_K$. While the proximal point algorithm is general, it is not very effective for problems in high-dimensional machine learning that involve minimizing a {\em sum} of two or more functions, one or more of which may not be differentiable. A key extension of the proximal point algorithm is through a general decomposition principle called operator splitting, reviewed below.
\section{Decomposition through Operator Splitting}
Operator splitting \cite{operator-splitting,douglas-rachford} is a generic approach to decomposing complex optimization and variational inequality problems into simpler ones that involve computing the resolvents of individual relations, rather than sums or other compositions of relations. For example, given a monotone inclusion problem of the form:
\[ 0 \in A(x) + B(x) \]
for two relations $A$ and $B$, how can we find the solution $x^*$ without computing the resolvent $(I + \lambda (A + B))^{-1}$, which may be complicated, but rather only compute the resolvents of $A$ and $B$ individually? There are several classes of operator splitting schemes. We will primarily focus on the {\em Douglas Rachford} algorithm \cite{douglas-rachford} specified in Figure~\ref{splitting}, because it leads to a widely used distributed optimization method called Alternating Direction Method of Multipliers (ADMM) \cite{boyd:admm}. The Douglas Rachford method is based on the ``damped iteration" given by:
\[ z_{k+1} = \frac{1}{2} (I + C_A C_B) (z_k) \]
where $C_A = 2 R_A + I$ and $C_B = 2 R_B + I$ are the ``reflection" or {\em Cayley} operators associated with the relations $A$ and $B$. Note that the Cayley operator is defined in terms of the resolvent, so this achieves the necessary decomposition.
\begin{figure} [th]
\centering
\fbox{
\begin{minipage}[t]{0.4\textwidth}
\begin{algorithm}[H]
\caption{Douglas Rachford method.}
{\bf INPUT:} Given $A(x), B(X)$ and a scalar $\lambda > 0$.
\begin{algorithmic}[1]
\STATE Set $k=0$ and initial vector $z_k = 0$.
\REPEAT
\STATE Set $x_{k + \frac{1}{2}} \leftarrow R_B(z_k)$
\STATE Set $z_{k+\frac{1}{2}} \leftarrow 2 x_{k + \frac{1}{2}} - z_k$
\STATE Set $x_{k+1} \leftarrow R_A(z_{k + \frac{1}{2}})$
\STATE Set $z_{k+1} \leftarrow z_k + x_{k+1} - x_{k+ \frac{1}{2}}$
\UNTIL{$z_{k + 1} < \epsilon$ }
\STATE Return $x_{k+1}$
\end{algorithmic}
\end{algorithm}
\end{minipage}
\begin{minipage}[t]{0.55\textwidth}
\begin{algorithm}[H]
\caption{Alternating Direction Method of Multipliers.}
{\bf INPUT:} Given sub-differentiable convex functions $f(x), g(x)$ and a scalar $\lambda > 0$.
\begin{algorithmic}[1]
\STATE Set $k=0$ and initial vector $z_k = 0$.
\REPEAT
\STATE Set $x_{k + \frac{1}{2}} \leftarrow \mbox{argmin}_x (f(x) + \frac{1}{2 \lambda} \|x - z_k \|^2_2)$
\STATE Set $z_{k+\frac{1}{2}} \leftarrow 2 x_{k + \frac{1}{2}} - z_k$
\STATE Set $x_{k+1} \leftarrow \mbox{argmin}_x (g(x) + \frac{1}{2 \lambda} \|x - x_{k + \frac{1}{2}} \|^2_2)$
\STATE Set $z_{k+1} \leftarrow z_k + x_{k+1} - x_{k+ \frac{1}{2}}$
\UNTIL{$z_{k + 1} < \epsilon$ }
\STATE Return $x_{k+1}$
\end{algorithmic}
\end{algorithm}
\end{minipage}}
\caption{Operator splitting is a generic framework for decomposing a composite objective function into simpler components.}
\label{splitting}
\end{figure}
When $A = \partial f$ and $B = \partial g$, two convex functions, the Douglas Rachford algorithm becomes the well-known Alternating Direction Method of Multipliers (ADMM) method, as described in Figure~\ref{splitting}, where the resolvent of $A$ and $B$ turn into proximal minimization steps. The ADMM algorithm has been extensively studied in optimization; a detailed review is available in the tutorial paper by Boyd and colleagues \cite{boyd:admm}, covering both its theoretical properties, operator splitting origins, and applications to high-dimensional data mining. ADMMs have also recently been studied for spectroscopic data, in particular {\em hyperspectral unmixing} \cite{admm:hyper}.
\subsection{Forward Backwards Splitting}
In this section we will give a brief overview of proximal splitting
algorithms \cite{PROXSPLITTING2011}. The two key ingredients of proximal
splitting are proximal operators and operator splitting. Proximal
methods \cite{rockafellar1976PROXIMALmonotone,parikh2013proximal},
which are widely used in machine learning, signal processing, and stochastic
optimization, provide a general framework for large-scale optimization.
The {\em proximal mapping} associated with a convex function $h$ is defined
as:
\begin{equation}
{\rm{prox}}_{h}(x)=\arg\mathop{\min}\limits _{u}(h(u)+\frac{1}{2}{\left\Vert {u-x}\right\Vert ^{2}})\label{eq:prox}
\end{equation}
Operator splitting is widely used to reduce the computational complexity
of many optimization problems, resulting in algorithms such as
sequential non-iterative approach (SNIA), Strang splitting, and sequential
iterative approach (SIA). Proximal splitting is a technique that combines
proximal operators and operator splitting, and deals with
problems where the proximal operator is difficult to compute at first,
yet is easier to compute after decomposition. The very basic scenario
is Forward-Backward Splitting (FOBOS) \cite{FOBOS:2009}
\begin{equation}
\mathop{\min}\limits _{\theta}\left({\Psi(\theta)=f(\theta)+h(\theta)}\right)
\label{ch1eq:fobos}
\end{equation}
where $f(x)$ is a convex, continuously differentiable function with
$L$-Lipschitz-continuous bounded gradients, i.e. $\forall x,y,||\nabla f(x)-\nabla f(y)||\le L||x-y||$,
and $h(\theta)$ is a convex (possibly not smooth) function. FOBOS
solves this problem via the following proximal gradient method
\begin{equation}
{\theta _{t + 1}} = {\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}({\theta _t} - {\alpha _t}\nabla f({\theta _t}))
\label{eq:fobos-prox}
\end{equation}
An extension of FOBOS is when the objective function is separable,
i.e.,
\begin{equation}
\mathop{\min}\limits _{\theta}\sum\limits _{i=1}^{m}{{f_{i}}(\theta)}\label{eq:admm-probform}
\end{equation}
where computing ${\rm{pro}}{{\rm{x}}_{\sum\limits_{i = 1}^m {{f_i}} }}(\cdot)$
is difficult, yet for each $i$, ${\rm{pro}}{{\rm{x}}_{{f_i}}}(\cdot)$ is easy to
compute. To solve this problem, Douglas-Rachford splitting \cite{PROXSPLITTING2011}
and Alternating Direction of Multiple Multipliers (ADMM) can be used.
Recently, ADMM has been used proposed for sparse RL \cite{ZHIWEI2014}.
\subsection{Nonlinear Primal Problem Formulation}
In this paper we will investigate a scenario of proximal splitting that is
different from the problem formulation in Section (\ref{eq:admm-probform}),
namely the nonlinear primal form
\begin{equation}
\mathop{\min}\limits _{\theta}\left(\Psi(\theta)={F(K(\theta))+h(\theta)}\right)\label{eq:primal}
\end{equation}
where $F(\cdot)$ is a lower-semicontinuous (l.s.c) nonlinear convex
function, $K$ is a linear operator, the induced
norm is $||K||$. In the following, we will denote $F(K(\theta))$
as $F\circ K(\theta)$. The proximal operator of this problem is
\begin{equation}
{\theta _{t + 1}} = \arg \mathop {\min }\limits_\theta \{ \Psi (\theta ) + \frac{1}{{2{\alpha _t}}}||\theta - {\theta _t}||_2^2\} = {\rm{pro}}{{\rm{x}}_{{\alpha _t}}}_{\left( {F \circ K + h} \right)}({\theta _t})
\end{equation}
In many cases, although ${{\rm{pro}}{{\rm{x}}_{{\alpha _t}F}}}$ and ${{\rm{pro}}{{\rm{x}}_{{\alpha _t}K}}}$
are easy to compute, ${{\rm{pro}}{{\rm{x}}_{{\alpha _t}F\circ K}}}$ is often difficult
to compute. For the NEU case, we have
\begin{equation}
\begin{array}{l}
K(\theta)=\mathbb{E}[{\phi_{t}}{\delta_{t}}]={\Phi^{T}}\Xi(T{V_{\theta}}-{V_{\theta}})={\Phi^{T}}\Xi(R+\gamma{\Phi^{'}}\theta-\Phi\theta),{\rm {}}F(\cdot)=\frac{1}{2}||\cdot||_{2}^{2}\end{array}\label{eq:neu-fk}
\end{equation}
It is straightforward to verify that ${{\rm{pro}}{{\rm{x}}_{{\alpha _t}F}}},{{\rm{pro}}{{\rm{x}}_{{\alpha _t}K}}}$
are easy to compute, but ${{\rm{pro}}{{\rm{x}}_{{\alpha _t}F\circ K}}}$
is not easy to compute since it involves the biased sampling problem
as indicated in Equation (\ref{eq:neu-grad}). To solve this problem,
we transform the problems formulation to facilitate operator splitting,
i.e., which only uses ${{\rm{pro}}{{\rm{x}}_{{\alpha _t}F}}},{{\rm{pro}}{{\rm{x}}_{{\alpha _t}K},}}{{\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}}$
and avoids computing ${{\rm{pro}}{{\rm{x}}_{{\alpha _t}F\circ K}}}$ directly. We
will use the primal-dual splitting framework to this end.
\subsection{Primal-Dual Splitting}
\label{pd-splitting}
The corresponding primal-dual formulation \cite{BOOK2011PROXSPLIT,PROXSPLITTING2011,POCK2011SADDLE}
of Section (\ref{eq:primal}) is
\begin{equation}
\mathop{\min}\limits _{\theta\in X}\mathop{\max}\limits _{y\in Y}\left({L(\theta,y)=\left\langle {K(\theta),y}\right\rangle -{F^{*}}(y)}+h(\theta)\right)
\label{eq:primal-dual}
\end{equation}
where $F^{*}(\cdot)$ is the Legendre transform of the convex nonlinear
function $F(\cdot)$, which is defined as $\ensuremath{{F^{*}}(y)={\sup_{x\in X}}(\langle x,y \rangle -F(x))}$.
The proximal splitting update per iteration is written as
\begin{equation}
\begin{array}{l}
{y_{t + 1}} = \arg \mathop {\min }\limits_{y \in Y} \left\langle { - {K_t}({\theta _t}),y} \right\rangle + {F^*}(y) + \frac{1}{{2{\alpha _t}}}||y - {y_t}|{|^2}\\
{{\theta_{t+1}}=\arg\mathop{\min}\limits _{\theta\in X}\left\langle {{K_{t}}({\theta}),y_{t}}\right\rangle +h(\theta)+\frac{1}{{2{\alpha_{t}}}}||\theta-{\theta_{t}}|{|^{2}}}
\end{array}
\end{equation}
Thus we have the general update rule as
\begin{equation}
\begin{array}{*{20}{l}}
{{y_{t + 1}} = {y_t} + {\alpha _t}{K_t}({\theta _t}) - {\alpha _t}\nabla F_t^*(y)
{\;,\;}
{\theta _{t + 1}} = {\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}({\theta _t} - {\alpha _t}\nabla {K_t}({\theta _t}){y_t})}
\end{array}
\end{equation}
However, in stochastic learning setting, we do not have knowledge of the exact ${K_t}({\theta _t})$, $\nabla F_t^*(y)$ and $\nabla {K_t}({\theta _t}){y_t}$, whereas a stochastic oracle $\mathcal{SO}$ is able to provide unbiased estimation of them.
\section{Natural Gradient Methods}
\label{sec:NG}
Consider the problem of minimizing a differentiable function $f:\mathbb R^n \to \mathbb R$. The standard gradient descent approach is to select an initial $x_0 \in \mathbb R^n$, compute the direction of steepest descent, $-\nabla f(x_0)$, and then move some amount in that direction (scaled by a stepsize parameter, $\alpha_0$). This process is then repeated indefinitely: $x_{k+1} = x_k - \alpha_k \nabla f(x_k)$, where $\{\alpha_k\}$ is a stepsize schedule and $k \in \{1,\hdots\}$. Gradient descent has been criticized for its low asymptotic rate of convergence. Natural gradients are a quasi-Newton approach to improving the convergence rate of gradient descent.
When computing the direction of steepest descent, gradient descent assumes that the vector $x_k$ resides in Euclidean space. However, in several settings it is more appropriate to assume that $x_k$ resides in a Riemannian space with metric tensor $G(x_k)$, which is an $n \times n$ positive definite matrix that may vary with $x_k$ \cite{Amari1998b}. In this case, the direction of steepest descent is called the \emph{natural gradient} and is given by $-G(x_k)^{-1}\nabla f(x_k)$ \cite{Amari1998}. In certain cases, (which include our policy search application), following the natural gradient is asymptotically Fisher-efficient \cite{Amari1998b}.
\section{Summary}
We provided a brief overview of some background material in reinforcement learning and optimization in this chapter. The subsequent chapters contain further elaboration of this material as it is required.
The overall goal of our work is to bring reinforcement learning into the main fabric of modern stochastic optimization theory. As we show in subsequent chapters, accomplishing this goal gives us access
to many advanced algorithms and analytical tools. It is worth noting that we make little use of classical stochastic approximation theory, which has traditionally been used to analyze reinforcement learning
methods (as discussed in detail in books such as \cite{ndp:book}). Classical stochastic approximation theory provides only asymptotic convergence bounds, for the most part. We are interested, however, in
getting tighter sample complexity bounds, which stochastic optimization provides.
\chapter{Sparse Temporal Difference Learning in Primal Dual Spaces}
\label{md-td-chapter}
In this chapter we explore a new framework for (on-policy convergent)
TD learning algorithm based on \textit{mirror descent} and related
algorithms.\footnote{This chapter is based on the paper ``Sparse Q-learning with Mirror Descent" published in UAI 2012.}
Mirror descent can be viewed as an enhanced gradient method,
particularly suited to minimization of convex functions in high-dimensional
spaces. Unlike traditional gradient methods, mirror descent undertakes
gradient updates of weights in both the dual space and primal space,
which are linked together using a Legendre transform. Mirror descent
can be viewed as a proximal algorithm where the distance-generating
function used is a Bregman divergence. A new class of {\em proximal-gradient}
based temporal-difference (TD) methods are presented based on different
Bregman divergences, which are more powerful than regular TD learning.
Examples of Bregman divergences that are studied include $p$-norm
functions, and Mahalanobis distance based on the covariance of sample
gradients. A new family of sparse mirror-descent reinforcement learning
methods are proposed, which are able to find sparse fixed-point of
an $l_{1}$-regularized Bellman equation at significantly less computational
cost than previous methods based on second-order matrix methods.
\section{Problem Formulation}
The problem formulation in this chapter is based on the Lasso-TD objective
defined as follows,
which is used in LARS-TD and LCP-TD. We first define $l_{1}$-regularized Projection, and then give the definition of Lasso-TD objective function.
\begin{defn}
\cite{LASSOTD:2011} ($l_{1}$-regularized Projection): ${{\Pi}_{{l_{1}}}}$ is the $l_{1}$-regularized
projection defined as:
\[{\Pi_{{l_{1}}}}y=\Phi\theta, \theta=\arg{\min_{w}}{\left\Vert {y-\Phi w}\right\Vert ^{2}}+\rho{\left\Vert w\right\Vert _{1}}\]
which is a non-expansive mapping w.r.t weighted $l_{2}$ norm, as proven in \cite{LASSOTD:2011}.
\end{defn}
\begin{lemma}
\cite{LASSOTD:2011}: ${\Pi_{\rho}}$ is a non-expansive mapping such
that
\begin{equation}
\forall x,y\in{R^{d}},||{\Pi_{\rho}}x-{\Pi_{\rho}}y|{|^{2}}\le||x-y|{|^{2}}-||x-y-({\Pi_{\rho}}x-{\Pi_{\rho}}y)|{|^{2}}\label{eq:proj_rho}
\end{equation}
\end{lemma}
\begin{defn}
\cite{LASSOTD:2011} (Lasso-TD) Lasso-TD is a fixed-point
equation w.r.t $l_{1}$ regularization with
parameter $\rho$, which is defined as
\begin{equation}
\begin{array}{l}
\theta=f(\theta)={\rm {argmi}}{{\rm {n}}_{u\in{R^{d}}}}\left({||T\Phi\theta-\Phi u{||{}^{2}}+\rho||u{||_{1}}}\right)\\
{\rm {=argmi}}{{\rm {n}}_{u\in{R^{d}}}}\left({||{R^{\pi}}+\gamma{P^{\pi}}\Phi\theta-\Phi u{||{}^{2}}+\rho||u{||{}_{1}}}\right)
\end{array}\label{eq:lassotd}
\end{equation}
\end{defn}
The properties of Lasso-TD is discussed in detail in \cite{LASSOTD:2011}.
Note that the above $l_{1}$ regularized fixed-point is not a convex
optimization problem but a fixed-point problem. Several prevailing
sparse RL methods use Lasso-TD as the objective function, such as SparseTD\cite{mahadevan:uai2005},
LARS-TD\cite{Kolter09LARSTD} and LCP-TD\cite{JeffLcp:nips2010}.
The advantage of LARS-TD comes from LARS in that it computes a homotopy path of solutions with
different regularization parameters, and thus offers a rich solution
family. The major drawback comes from LARS, too. To maintain the LARS
criteria wherein each active variable has the same correlation with
the residual, variables may be added and dropped several times, which
is computationally expensive. In fact, the computational complexity
per iteration is $O(Ndk^{2})$ where $k$ is the cardinality of the active
feature set. Secondly, LARS-TD requires the $A$ matrix to be a $P$-matrix(a
square matrix which does not necessarily to be symmetric, but all
the principal minors are positive), which poses extra limitation on
applications. The author of LARS-TD claims that this seems never to
be a problem in practice, and given on-policy sampling condition or
given large enough ridge regression term, $P$-matrix condition can
be guaranteed. LCP-TD {[}12{]} formulates LASSO-TD as a linear complementarity
problem (LCP), which can be solved by a variety of available LCP solvers.
We then derive the major step by formulating the
problem as a forward-backward splitting problem (FOBOS) as in \cite{FOBOS:2009},
\begin{equation}
\begin{array}{l}
{\theta{}_{t+\frac{1}{2}}}={\theta_{t}}-{\alpha_{t}}{g_{t}}\\
{\theta_{t+1}}=\arg\mathop{\min}\limits _{\theta}\left\{ {\frac{1}{2}||{\theta}-{\theta_{t+\frac{1}{2}}}||_{2}^{2}+\alpha_{t}h(\theta)}\right\}
\end{array}\label{eq:fobos}
\end{equation}
This is equivalent to the formulation of proximal gradient method
\begin{equation}
{\theta_{t+1}}=\arg\mathop{\min}\limits _{\theta}\left\{ {\left\langle {{g_{t}},\theta}\right\rangle +h(\theta)+\frac{1}{2\alpha_{t}}||\theta-{\theta_{t}}||_{2}^{2}}\right\} \label{eq:fobos2}
\end{equation}
Likewise, we could formulate the sparse TD algorithm as
\begin{equation}
\begin{array}{l}
{\theta_{t+\frac{1}{2}}}={\theta_{t}}-\frac{{\alpha_{t}}}{2}\nabla{\rm {MSE(}}\theta{\rm {)}}\\
{\theta_{t+1}}=\arg\mathop{\min}\limits _{\theta}\left\{ {\frac{1}{2}||{\theta_{t}}-{\theta_{t+\frac{1}{2}}}||_{2}^{2}+\alpha_{t}h(\theta)}\right\}
\end{array}
\end{equation}
And this can be formulated as
\begin{equation}
{\theta_{t+1}}=\arg\mathop{\min}\limits _{\theta}\left\{ {\left\langle {\frac{1}{2}\nabla{\rm {MSE(}}\theta{\rm {)}},\theta}\right\rangle +h(\theta)+\frac{1}{2\alpha_{t}}||\theta-{\theta_{t}}||_{2}^{2}}\right\}
\end{equation}
\section{Mirror Descent RL}
\begin{algorithm}
\label{mdtd-algm1}
\caption{Adaptive Mirror Descent TD($\lambda$)}
Let $\pi$ be some fixed policy for an MDP M, and $s_0$ be the initial state. Let $\Phi$ be some fixed or automatically generated basis.
\begin{algorithmic}[1]
\REPEAT
\STATE Do action $\pi(s_t)$ and observe next state $s_{t+1}$ and reward $r_t$.
\STATE Update the eligibility trace $e_t \leftarrow e_t + \lambda \gamma \phi(s_t)$
\STATE Update the dual weights $\theta_t$ for a linear function approximator:
\[ \theta_{t+1} = \nabla \psi_t(w_t) + \alpha_t (r_t + \gamma \phi(s_{t+1})^T w_t - \phi(s_t)^T w_t) e_t \]
where $\psi$ is a distance generating function.
\STATE Set $w_{t+1} = \nabla \psi_t^*(\theta_{t+1})$ where $\psi^*$ is the Legendre transform of $\psi$.
\STATE Set $t \leftarrow t+1$.
\UNTIL{{\bf done}}.
Return $\hat{V}^\pi \approx \Phi w_t $ as the value function associated with policy $\pi$ for MDP $M$.
\end{algorithmic}
\end{algorithm}
Algorithm 1 describes the proposed mirror-descent TD($\lambda$) method.\footnote{All the algorithms described extend to the action-value case where $\phi(s)$ is replaced by $\phi(s,a)$. } Unlike regular TD, the weights are updated using the TD error in the dual space by mapping the primal weights $w$ using a gradient of a strongly convex function $\psi$. Subsequently, the updated dual weights are converted back into the primal space using the gradient of the Legendre transform of $\psi$, namely $\nabla \psi^*$. Algorithm 1 specifies the mirror descent TD($\lambda$) algorithm wherein each weight $w_i$ is associated with an eligibility trace $e(i)$. For $\lambda = 0$, this is just the features of the current state $\phi(s_t)$, but for nonzero $\lambda$, this corresponds to a decayed set of features proportional to the recency of state visitations. Note that the distance generating function $\psi_t$ is a function of time.
\subsection{Choice of Bregman Divergence}\label{bd}
We now discuss various choices for the distance generating function in Algorithm 1. In the simplest case, suppose $\psi(w) = \frac{1}{2} \| w \|^2_2$, the Euclidean length of $w$. In this case, it is easy to see that mirror descent TD($\lambda$) corresponds to regular TD($\lambda$), since the gradients $\nabla \psi$ and $\nabla \psi^*$ correspond to the identity function.
A much more interesting choice of $\psi$ is $\psi(w) = \frac{1}{2} \| w \|^2_q$, and its conjugate Legendre transform $\psi^*(w) = \frac{1}{2} \| w \|^2_p$. Here, $\|w\|_q = \left( \sum_j |w_j|^q \right)^{\frac{1}{q}}$, and $p$ and $q$ are conjugate numbers such that $\frac{1}{p} + \frac{1}{q} = 1$.
This $\psi(w)$ leads to the p-norm link function $\theta = f(w)$ where $f: \mathbb{R}^d \rightarrow \mathbb{R}^d$ \cite{gentile:mlj}:
\begin{equation}
\label{pnorm}
f_j(w) = \frac{\mbox{sign}(w_j) |w_j|^{q-1}}{\| w \|^{q-2}_q}, \ \ f^{-1}_j(\theta) = \frac{\mbox{sign}(\theta_j) |\theta_j|^{p-1}}{\| \theta \|^{p-2}_p}
\end{equation}
The p-norm function has been extensively studied in the literature on online learning \cite{gentile:mlj}, and it is well-known that for large $p$, the corresponding classification or regression method behaves like a multiplicative method (e.g., the p-norm regression method for large $p$ behaves like an exponentiated gradient method (EG) \cite{kivinen:95,littlestone:mlj}).
Another distance generating function is the negative entropy function $\psi(w) = \sum_i w_i \log w_i$, which leads to the entropic mirror descent algorithm \cite{Beck:2003p2359}. Interestingly, this special case has been previously explored \cite{sutton-precup:egtd} as the exponentiated-gradient TD method, although the connection to mirror descent and Bregman divergences were not made in this previous study, and EG does not generate sparse solutions \cite{ShalevShwartz:jmlr}. We discuss EG methods vs. p-norm methods in Section~\ref{eg-vs-pnorm}.
\subsection{Sparse Learning with Mirror Descent TD}\label{l1mdtd}
\begin{algorithm}
\caption{Sparse Mirror Descent TD($\lambda$)}
\begin{algorithmic}[1]
\REPEAT
\STATE Do action $\pi(s_t)$ and observe next state $s_{t+1}$ and reward $r_t$.
\STATE Update the eligibility trace $e_t \leftarrow e_t + \lambda \gamma \phi(s_t)$
\STATE Update the dual weights $\theta_t$:
\[ \tilde{\theta}_{t+1} = \nabla \psi_t(w_t) + \alpha_t \left( r_t + \gamma \phi(s_{t+1})^T w_t - \phi(s_t)^T w_t \right) e_t \]
(e.g., $\psi(w) = \frac{1}{2} \| w \|^2_q$ is the p-norm link function).
\STATE Truncate weights:
\[ \forall j, \ \ \theta^{t+1}_j = \mbox{sign}(\tilde{\theta}^{t+1}_j) \max(0, |\tilde{\theta}^{t+1}_j| - \alpha_t \beta) \]
\STATE $w_{t+1} = \nabla \psi_t^*(\theta_{t+1})$ (e.g., $\psi^*(\theta) = \frac{1}{2} \|\theta\|^2_p$ and $p$ and $q$ are dual norms such that $\frac{1}{p} + \frac{1}{q} = 1$).
\STATE Set $t \leftarrow t+1$.
\UNTIL{{\bf done}}.
Return $\hat{V}^\pi \approx \Phi w_t $ as the $l_1$ penalized sparse value function associated with policy $\pi$ for MDP $M$.
\end{algorithmic}
\end{algorithm}
Algorithm 2 describes a modification to obtain sparse value functions resulting in a sparse mirror-descent TD($\lambda)$ algorithm. The main difference is that the dual weights $\theta$ are truncated according to Equation~\ref{l1prox} to satisfy the $l_1$ penalty on the weights. Here, $\beta$ is a sparsity parameter. An analogous approach was suggested in \cite{ShalevShwartz:jmlr} for $l_1$ penalized classification and regression.
\subsection{Composite Mirror Descent TD}\label{comdtd}
Another possible mirror-descent TD algorithm uses as the distance-generating function a Mahalanobis distance derived from the subgradients generated during actual trials. We base our derivation on the composite mirror-descent approach proposed in \cite{duchi:jmlr} for classification and regression. The composite mirror-descent solves the following optimization problem at each step:
\begin{equation}
\label{adaptsubgmd}
w_{t+1} = \mbox{argmin}_{x \in X} \left( \alpha_t \langle x, \partial f_t \rangle + \alpha_t \mu(x) + D_{\psi_t}(x, w_t) \right)
\end{equation}
Here, $\mu$ serves as a fixed regularization function, such as the $l_1$ penalty, and $\psi_t$ is the time-dependent distance generating function as in mirror descent. We now describe a different Bregman divergence to be used as the distance generating function in this method. Given a positive definite matrix $A$, the Mahalanobis norm of a vector $x$ is defined as $\|x\|_A = \sqrt{\langle x, A x \rangle}$. Let $g_t = \partial f(s_t)$ be the subgradient of the function being minimized at time $t$, and $G_t = \sum_t g_t g_t^T$ be the covariance matrix of outer products of the subgradients. It is computationally more efficient to use the diagonal matrix $H_t = \sqrt{\mbox{diag}(G_t)}$ instead of the full covariance matrix, which can be expensive to estimate. Algorithm 3 describes the adaptive subgradient mirror descent TD method.
\begin{algorithm}
\caption{Composite Mirror Descent TD($\lambda$)}
\begin{algorithmic}[1]
\REPEAT
\STATE Do action $\pi(s_t)$ and observe next state $s_{t+1}$ and reward $r_t$.
\STATE Set TD error $\delta_t = r_t + \gamma \phi(s_{t+1})^T w_t - \phi(s_t)^T w_t$
\STATE Update the eligibility trace $e_t \leftarrow e_t + \lambda \gamma \phi(s_t)$
\STATE Compute TD update $\xi_t = \delta_t e_t$.
\STATE Update feature covariance
\[G_{t} = G_{t-1} + \phi(s_t) \phi(s_t)^T\]
\STATE Compute Mahalanobis matrix $H_{t} = \sqrt{\mbox{diag}(G_{t})}$.
\STATE Update the weights $w$:
\[ w_{{t+1},i} = \mbox{sign}( w_{t,i} - \frac{\alpha_t \xi_{t,i}}{H_{{t},ii}})(|w_{t,i} - \frac{\alpha_t \xi_{t,i} }{H_{{t},ii}}| - \frac{ \alpha_t \beta}{H_{{t},ii}} )\]
\STATE Set $t \leftarrow t+1$.
\UNTIL{{\bf done}}.
Return $\hat{V}^\pi \approx \Phi w_t $ as the $l_1$ penalized sparse value function associated with policy $\pi$ for MDP $M$.
\end{algorithmic}
\end{algorithm}
\section{Convergence Analysis}
\textbf{Definition 2} \cite{LASSOTD:2011}: ${{\Pi}_{{l_{1}}}}$
is the $l_{1}$-regularized projection defined as: ${\Pi_{{l_{1}}}}y=\Phi\alpha$ such that $\alpha=\arg{\min_{w}}{\left\Vert {y-\Phi w}\right\Vert ^{2}}+\beta{\left\Vert w\right\Vert _{1}}$, which is a non-expansive mapping w.r.t weighted $l_2$ norm induced by the on-policy sample distribution setting, as proven in \cite{LASSOTD:2011}. Let the approximation error $f(y,\beta)={\left\Vert {y-{\Pi_{{l_{1}}}}y}\right\Vert ^{2}}$.
\noindent \textbf{Definition 3} (Empirical $l_{1}$-regularized projection): ${{\hat{\Pi}}_{{l_{1}}}}$
is the empirical $l_{1}$-regularized projection with a specific $l_{1}$
regularization solver, and satisfies the non-expansive mapping property.
It can be shown using a direct derivation that ${{{\hat \Pi }_{{l_1}}}\Pi T}$ is a $\gamma$-contraction mapping. Any unbiased $l_1$ solver which generates intermediate sparse solution before convergence, e.g., SMIDAS solver after $t$-th iteration, comprises an empirical $l_{1}$-regularized projection.
\noindent \textbf{Theorem 1} The approximation error $||V - \hat V|| $ of Algorithm 2
is bounded by (ignoring dependence on $\pi$ for simplicity):
\begin{equation}
\begin{array}{*{20}{l}}
{|| {V - \hat V} || \le \frac{1}{{1 - \gamma }} \times }\\
{\left( {\left\| {V - \Pi V} \right\| + f(\Pi V,\beta ) + (M - 1)P(0) + \left\| {{w^*}} \right\|_1^2\frac{M}{{\alpha_t N}}} \right)}
\end{array}
\label{eq:theorem}
\end{equation}
where $\hat{V}$ is the approximated value function after $N$-th iteration, i.e., $\hat{V}=\Phi{w_{N}}$,
$M=\frac{2}{{2-4\alpha_t(p-1)e}}$, $\alpha_t$ is the stepsize, $P(0) = \frac{1}{N}\sum\limits_{i = 1}^N {\left\| {\Pi V({s_i})} \right\|_2^2} $, $s_i$ is the state of $i$-th sample, $e = {d^{\frac{p}{2}}}$, $d$ is the number of features, and finally, $w^*$ is $l_1$-regularized projection of $\Pi V$ such that $\Phi {w^*} = {\Pi _{{l_1}}}\Pi V$.
\noindent \textbf{Proof:} In the on-policy setting, the solution given by Algorithm 2 is the fixed point
of $\hat V = {\hat \Pi _{{l_1}}}\Pi T\hat V$
and the error decomposition is illustrated in Figure \ref{Fio:Error2}.
\begin{figure}
\centering
\includegraphics[width=2in,height=1.5in]{Error2.pdf}
\caption{Error Bound and Decomposition}
\label{Fio:Error2}
\end{figure}
The error can be bounded by the triangle inequality
\begin{small}
\begin{equation}
|| {V - \hat V} || = || {V - \Pi TV} || + ||\Pi TV - {\hat \Pi _{{l_1}}}\Pi TV|| + ||{\hat \Pi _{{l_1}}}\Pi TV - \hat V||
\label{eq:triangular}
\end{equation}
\end{small}
Since ${{{\hat \Pi }_{{l_1}}}\Pi T}$ is a $\gamma$-contraction mapping,
and $\hat V = {\hat \Pi _{{l_1}}}\Pi T\hat V$, we have
\begin{equation}
||{\hat \Pi _{{l_1}}}\Pi TV - \hat V|| = ||{\hat \Pi _{{l_1}}}\Pi TV - {\hat \Pi _{{l_1}}}\Pi T\hat V|| \le \gamma ||V - \hat V||
\label{eq:gamma}\end{equation}
So we have
\[(1 - \gamma )|| {V - \hat V} || \le || {V - \Pi TV} || + ||\Pi TV - {\hat \Pi _{{l_1}}}\Pi TV||\]
$\left\Vert {V-\Pi TV}\right\Vert $ depends on the expressiveness
of the basis $\Phi$, where if $V$ lies in $span(\Phi)$, this error term
is zero. $||\Pi TV - {\Pi _{{l_1}}}\hat \Pi TV||$
is further bounded by the triangle inequality
\[\begin{array}{*{20}{l}}
{||\Pi TV - {{\hat \Pi }_{{l_1}}}\Pi TV|| \le }\\
{||\Pi TV - {\Pi _{{l_1}}}\Pi TV|| + ||{\Pi _{{l_1}}}\Pi TV - {{\hat \Pi }_{{l_1}}}\Pi TV||}
\end{array}\]
where $\left\Vert {\Pi TV-{\Pi_{{l_{1}}}}\Pi TV}\right\Vert $ is controlled
by the sparsity parameter $\beta$, i.e.,
$f(\Pi TV,\beta ) = || {\Pi TV - {\Pi _{{l_1}}}\Pi TV} ||$,
where \noindent $\varepsilon = || {{{\hat \Pi }_{{l_1}}}\Pi TV - {\Pi _{{l_1}}}\Pi TV} ||$
is the approximation error depending on the quality of the $l_{1}$ solver employed. In Algorithm 2, the $l_{1}$
solver is related to the SMIDAS $l_1$ regularized mirror-descent method for regression and classification \cite{ShalevShwartz:jmlr}. Note that for a squared loss function
$L(\left\langle {w,{x_{i}}}\right\rangle ,{y_{i}})=||\left\langle {w,{x_{i}}}\right\rangle -{y_{i}}||_{2}^{2}$, we have ${\left|{L'}\right|^{2}}\le 4L$.
Employing the result of Theorem 3 in \cite{ShalevShwartz:jmlr}, after the $N$-th iteration, the $l_{1}$ approximation error is bounded by
\[\varepsilon \le (M - 1)P(0) + || {{w^*}} ||_1^2\frac{M}{{\alpha_t N}},M = \frac{2}{{2 - 4\alpha_t (p - 1)e}}\]
By rearranging the terms and applying $V=TV$, Equation (\ref{eq:theorem}) can be deduced.
\section{Experimental Results: Discrete MDPs}
Figure~\ref{ogrid-results} shows that mirror-descent TD converges more quickly with far smaller Bellman errors than LARS-TD \cite{lars-td} on a discrete ``two-room" MDP \cite{pvf-jmlr1}. The basis matrix $\Phi$ was automatically generated as $50$ proto-value functions by diagonalizing the graph Laplacian of the discrete state space connectivity graph\cite{pvf-jmlr1}. The figure also shows that Algorithm 2 (sparse mirror-descent TD) scales more gracefully than LARS-TD. Note LARS-TD is unstable for $\gamma = 0.9$.
It should be noted that the computation cost of LARS-TD is $O(Ndm^3)$, whereas that for Algorithm 2 is $O(Nd)$, where $N$ is the number of samples, $d$ is the number of basis functions, and $m$ is the number of active basis functions. If $p$ is linear or sublinear w.r.t $d$, Algorithm 2 has a significant advantage over LARS-TD.
\begin{figure}[ht]
\begin{center}
\begin{minipage}[t]{0.4\textwidth}
\includegraphics[height=0.15\textheight,width=2.8in]{ogrid-bdq-larstd-convergence-crop.pdf}\\
\includegraphics[height=0.15\textheight,width=2.8in]{ogrid-bdq-larstd-convergence-low-gamma-crop.pdf}\\
\includegraphics[height=0.15\textheight,width=2.8in]{ogrid-bdq-larstd-times-crop.pdf}
\end{minipage}
\end{center}
\caption{Mirror-descent Q-learning converges significantly faster than LARS-TD on a ``two-room" grid world MDP for $\gamma = 0.9$ (top left) and $\gamma = 0.8$ (top right). The y-axis measures the $l_2$ (red curve) and $l_\infty$ (blue curve) norm difference between successive weights during policy iteration. Bottom: running times for LARS-TD (blue solid) and mirror-descent Q (red dashed). Regularization $\beta = 0.01$. }
\label{ogrid-results}
\end{figure}
\begin{figure}[ht]
\centering
\begin{minipage}[t]{0.6\textwidth}
\includegraphics[width=3in]{noisy-pvfs.png}
\end{minipage} \hspace{0.2in}
\caption{Sensitivity of sparse mirror-descent TD to noisy features in a grid-world domain. Left: basis matrix with the first 50 columns representing proto-value function bases and the remainder 450 bases representing mean-0 Gaussian noise. Right: Approximated value function using sparse mirror-descent TD. }
\label{grid-noisypvfs}
\end{figure}
Figure~\ref{grid-noisypvfs} shows the result of another experiment conducted to test the noise immunity of Algorithm 2 using a discrete $10 \times 10$ grid world domain with the goal set at the upper left hand corner. For this problem, $50$ proto-value basis functions were automatically generated, and $450$ random Gaussian mean $0$ noise features were added. The sparse mirror descent TD algorithm was able to generate a very good approximation to the optimal value function despite the large number of irrelevant noisy features, and took a fraction of the time required by LARS-TD.
Figure~\ref{pnorm-decay} compares the performance of mirror-descent Q-learning with a fixed p-norm link function vs. a decaying p-norm link function for a $10 \times 10$ discrete grid world domain with the goal state in the upper left-hand corner. Initially, $p = O(\log d)$ where $d$ is the number of features, and subsequently $p$ is decayed to a minimum of $p=2$. Varying $p$-norm interpolates between additive and multiplicative updates. Different values of $p$ yield an interpolation between the truncated gradient method \cite{Lanford:2008} and SMIDAS \cite{SMIDAS:2009}.
\begin{figure}[ht]
\begin{center}
\begin{minipage}[t]{0.4\textwidth}
\includegraphics[totalheight=0.13\textheight,width=2.8in]{bdq-grid-no-decay.pdf}\\
\includegraphics[totalheight=0.13\textheight,width=2.8in]{bdq-grid-sqrt-decay-pnorm.pdf}
\end{minipage}
\end{center}
\caption{Left: convergence of mirror-descent Q-learning with a fixed p-norm link function. Right: decaying p-norm link function. }
\label{pnorm-decay}
\end{figure}
\begin{figure}[t]
\begin{minipage}[t]{0.2\textwidth}
\includegraphics[totalheight=0.13\textheight, width=2in]{comd-ogrid-pvf.pdf}
\end{minipage}
\begin{minipage}[t]{0.2\textwidth}
\includegraphics[totalheight=0.13\textheight, width=2in]{comd-ogrid-vf2-crop.pdf}
\end{minipage}
\caption{Left: Convergence of composite mirror-descent Q-learning on two-room gridworld domain. Right: Approximated value function, using $50$ proto-value function bases.}
\label{comd-ogrid-results}
\end{figure}
Figure~\ref{comd-ogrid-results} illustrates the performance of Algorithm 3 on the two-room discrete grid world navigation task.
\section{Experimental Results: Continuous MDPs}\label{contmdp}
Figure~\ref{mcar-experiment} compares the performance of Q-learning vs. mirror-descent Q-learning for the mountain car task, which converges more quickly to a better solution with much lower variance. Figure~\ref{acrobot-experiment} shows that mirror-descent Q-learning with learned diffusion wavelet bases converges quickly on the $4$-dimensional Acrobot task. We found in our experiments that LARS-TD did not converge within $20$ episodes (its curve, not shown in Figure~\ref{mcar-experiment}, would be flat on the vertical axis at $1000$ steps).
\begin{figure}[tbh]
\begin{center}
\begin{minipage}[t]{0.4\textwidth}
\includegraphics[height=1.5in, width=2in]{mcar-qlambda-25bases-crop.pdf}
\end{minipage} \hspace{.1in}
\begin{minipage}[t]{0.4\textwidth}
\includegraphics[height=1.5in, width=2.5in]{mcar-bdq-25bases-crop.pdf}
\end{minipage}
\end{center}
\caption{Top: Q-learning; Bottom: mirror-descent Q-learning with p-norm link function, both with $25$ fixed Fourier bases \cite{konidaris2008value} for the mountain car task. }
\label{mcar-experiment}
\end{figure}
\begin{figure}[tbh]
\begin{center}
\begin{minipage}[t]{0.45\textwidth}
\includegraphics[height=1.35in, width=2in]{acrobot-bdq-dwt-25bases-crop.pdf}
\end{minipage}
\end{center}
\caption{Mirror-descent Q-learning on the Acrobot task using automatically generated diffusion wavelet bases averaged over $5$ trials. }
\label{acrobot-experiment}
\end{figure}
Finally, we tested the mirror-descent approach on a more complex $8$-dimensional continuous MDP. The triple-link inverted pendulum \cite{si2001online} is a highly nonlinear time-variant under-actuated system, which is a standard benchmark testbed in the control community. We base our simulation using the system parameters described in \cite{si2001online}, except that the action space is discretized because the algorithms described here are restricted to policies with discrete actions. There are three actions, namely $\{ 0,5{\rm{Newton}}, - 5{\rm{Newton}}\} $. The state space is $8$-dimensional, consisting of the angles made to the horizontal of the three links in the arm as well as their angular velocities, the position and velocity of the cart used to balance the pendulum. The goal is to learn a policy that can balance the system with the minimum number of episodes. A run is successful if it balances the inverted pendulum for the specified number of steps within $300$ episodes, resulting in a reward of $0$. Otherwise, this run is considered as a failure and yields a negative reward $-1$. The first action is chosen randomly to push the pendulum away from initial state. Two experiments were conducted on the triple-link pendulum domain with $20$ runs
for each experiment. As Table 1 shows, Mirror Descent Q-learning is able to learn the policy with fewer episodes and usually with reduced variance compared with regular Q-learning.
The experiment settings are Experiment 1: Zero initial state and the system receives a reward $1$ if it is able to balance 10,000 steps.
Experiment 2: Zero initial state and the system receives a reward $1$ if it is able
to balance 100,000 steps.
Table 1 shows the comparison result between regular Q-learning and Mirror Descent
Q-learning.
\begin{table}
\centering
\begin{tabular}{|c|c|c|}
\hline
\# of Episodes\textbackslash{}Experiment & 1 & 2\tabularnewline
\hline
Q-learning & $6.1 \pm 5.67$ & $15.4 \pm 11.33$ \tabularnewline
\hline
Mirror Descent Q-learning & $5.7 \pm 9.70$ & $11.8 \pm 6.86$ \tabularnewline
\hline
\end{tabular}
\caption{Results on Triple-Link Inverted Pendulum Task.}
\label{table1}
\end{table}
\section{Comparison of Link Functions}\label{eg-vs-pnorm}
The two most widely used link functions in mirror descent are the $p$-norm link function \cite{Beck:2003p2359} and the relative entropy function for exponentiated gradient (EG) \cite{kivinen:95}. Both of these link functions offer a multiplicative update rule compared with regular additive gradient methods. The differences between these two are discussed here. Firstly, the loss function for EG is the relative entropy whereas that of the $p$-norm link function is the square $l_2$-norm function. Second and more importantly, EG does not produce sparse solutions since it must maintain the weights away from zero, or else its potential (the relative entropy) becomes unbounded at the boundary.
Another advantage of $p$-norm link functions over EG is that the $p$-norm link function offers a flexible interpolation between additive and multiplicative gradient updates. It has been shown that when the features are dense and the optimal coefficients $\theta^*$ are sparse, EG converges faster than the regular additive gradient methods \cite{kivinen:95}. However, according to our experience, a significant drawback of EG is the overflow of the coefficients due to the exponential operator. To prevent overflow, the most commonly used technique is rescaling: the weights are re-normalized to sum to a constant. However, it seems that this approach does not always work. It has been pointed out \cite{sutton-precup:egtd} that in the EG-Sarsa algorithm, rescaling can fail, and replacing eligible traces instead of regular additive eligible traces is used to prevent overflow. EG-Sarsa usually poses restrictions on the basis as well. Thanks to the flexible interpolation capability between multiplicative and additive gradient updates, the $p$-norm link function is more robust and applicable to various basis functions, such as polynomial, radial basis function (RBF), Fourier basis \cite{konidaris2008value}, proto-value functions (PVFs), etc.
\section{Summary}
We proposed a novel framework for reinforcement learning using mirror-descent online convex optimization. Mirror Descent Q-learning demonstrates the following advantage over regular Q learning: faster convergence rate and reduced variance due to larger stepsizes with theoretical convergence guarantees \cite{nemirovski:siam}. Compared with existing sparse reinforcement learning algorithms such as LARS-TD, Algorithm 2 has lower sample complexity and lower computation cost, advantages accrued from the first-order mirror descent framework combined with proximal mapping \cite{ShalevShwartz:jmlr}.
There are many promising future research topics along this direction.
We are currently exploring a mirror-descent fast-gradient RL method, which is both convergent off-policy and quicker than fast gradient TD methods such as GTD and TDC \cite{Sutton09fastgradient-descent}. To scale to large MDPs, we are investigating hierarchical mirror-descent RL methods, in particular extending SMDP Q-learning. We are also undertaking a more detailed theoretical analysis of the mirror-descent RL framework, building on existing analysis of mirror-descent methods \cite{duchi:jmlr,ShalevShwartz:jmlr}. Two types of theoretical investigations are being explored: regret bounds of mirror-descent TD methods, extending previous results \cite{warmuth:mlj} and convergence analysis combining robust stochastic approximation \cite{nemirovski:siam} and RL theory \cite{ndp:book,borkar:book}.
\chapter{Regularized Off-Policy Temporal Difference Learning}\label{chapter:Regularized-Off-Policy-TD}
In the last chapter we proposed an on-policy convergent sparse TD learning algorithm.
Although TD converges when samples are drawn ``on-policy''
by sampling from the Markov chain underlying a policy in a Markov
decision process (MDP), it can be shown to be divergent when samples
are drawn ``off-policy''.
In this chapter, the off-policy TD learning problem is formulated
from the stochastic optimization perspective. \footnote{This chapter is based on the paper "Regularized Off-Policy TD-Learning" published in NIPS 2012.} A novel objective function
is proposed based on the linear equation formulation of the TDC algorithm.
The optimization problem underlying off-policy TD methods, such as
TDC, is reformulated as a convex-concave saddle-point stochastic approximation
problem, which is both convex and incrementally solvable. A detailed
theoretical and experimental study of the RO-TD algorithm is presented.
\section{Introduction}
\subsection{Off-Policy Reinforcement Learning}
Off-policy learning refers to learning about one way of behaving,
called the \emph{target policy}, from sample sets that are generated
by another policy of choosing actions, which is called the \emph{behavior
policy}, or exploratory policy. As pointed out
in \cite{maei2011gradient}, the target policy is often a deterministic
policy that approximates the optimal policy, and the behavior policy
is often stochastic, exploring all possible actions in each state
as part of finding the optimal policy. Learning the target policy
from the samples generated by the behavior policy allows a greater
variety of exploration strategies to be used. It also enables learning
from training data generated by unrelated controllers, including manual
human control, and from previously collected data. Another reason
for interest in off-policy learning is that it enables learning about
multiple target policies (e.g., optimal policies for multiple sub-goals)
from a single exploratory policy generated by a single behavior policy,
which triggered an interesting research area termed as ``parallel
reinforcement learning''.
Besides, off-policy methods are of
wider applications since they are able to learn while executing an
exploratory policy, learn from demonstrations, and learn multiple
tasks in parallel \cite{OFFACTOR:2012}. Sutton et al. \cite{FastGradient:2009}
introduced convergent off-policy temporal difference learning algorithms,
such as TDC, whose computation time scales linearly with the number
of samples and the number of features. Recently, a linear off-policy
actor-critic algorithm based on the same framework was proposed in
\cite{OFFACTOR:2012}.
\subsection{Convex-concave Saddle-point First-order Algorithms \label{sec:saddle}}
The key novel contribution of this chapter is a convex-concave saddle-point
formulation for regularized off-policy TD learning. A convex-concave
saddle-point problem is formulated as follows. Let $x\in X,y\in Y$,
where $X,Y$ are both nonempty bounded closed convex sets, and $f(x):X\to\mathbb{R}$
be a convex function. If there exists a function $\varphi(\cdot,\cdot)$
such that $f(x)$ can be represented as $f(x):={\sup_{y\in Y}}\varphi(x,y)$,
then the pair $(\varphi,Y)$ is referred as the saddle-point representation
of $f$. The optimization problem of minimizing $f$ over $X$ is
converted into an equivalent convex-concave saddle-point problem $SadVal={\inf_{x\in X}}{\sup_{y\in Y}}\varphi(x,y)$
of $\varphi$ on $X\times Y$. If $f$ is non-smooth yet convex and
well structured, which is not suitable for many existing optimization
approaches requiring smoothness, its saddle-point representation $\varphi$
is often smooth and convex. Thus, convex-concave saddle-point problems
are, therefore, usually better suited for first-order methods \cite{sra2011optimization}.
A comprehensive overview on extending convex minimization to convex-concave
saddle-point problems with unified variational inequalities is presented
in \cite{ben2005non}.
As an example, consider $f(x)=||Ax-b|{|_{m}}$ which admits a bilinear
minimax representation
\begin{equation}
f(x):={\left\Vert {Ax-b}\right\Vert _{m}}={\max_{{{\left\Vert y\right\Vert }_{n}}\le1}}{y^{T}}(Ax-b)\label{eq:minimax}
\end{equation}
where $m,n$ are conjugate numbers. Using the approach in \cite{RobustSA:2009},
Equation (\ref{eq:minimax}) can be solved as
\begin{equation}
{x_{t+1}}={x_{t}}-{\alpha_{t}}{A^{T}}{y_{t}},{y_{t+1}}={\Pi_{n}}({y_{t}}+{\alpha_{t}}(A{x_{t}}-b))\label{eq:minimaxFOM1}
\end{equation}
where $\Pi_{n}$ is the projection operator of $y$ onto the unit
$l_{n}$-ball ${\left\Vert y\right\Vert _{n}}\le{\rm {1}}$,which
is defined as
\begin{equation}
{\Pi_{n}}(y)=\min(1,1/{\left\Vert y\right\Vert _{n}})y,n=2,3,\cdots,{\Pi_{\infty}}{\rm {}}({y_{i}})=\min(1,1/|{y_{i}}|){y_{i}}\label{l2proj}
\end{equation}
and $\Pi_{\infty}$ is an entrywise operator.
\section{Problem Formulation}
\subsection{Objective Function Formulation}
Now let's review the concept of MSPBE. MSPBE is defined as
\begin{equation}
\begin{array}{l}
{\rm {MSPBE}}(\theta)\\
=\left\Vert {\Phi\theta-\Pi T(\Phi\theta)}\right\Vert _{\Xi}^{2}\\
={({\Phi^{T}}\Xi(T\Phi\theta-\Phi\theta))^{T}}{({\Phi^{T}}\Xi\Phi)^{-1}}{\Phi^{T}}\Xi(T\Phi\theta-\Phi\theta)\\
=\mathbb{E}{[\delta_t(\theta)\phi_t]^{T}}\mathbb{E}{[\phi_t{\phi_t^{T}}]^{-1}}\mathbb{E}[\delta_{t}(\theta)\phi_t]
\end{array}
\label{mspbe}
\end{equation}
To avoid computing the inverse matrix ${({\Phi^{T}}\Xi\Phi)^{-1}}$
and to avoid the double sampling problem \cite{sutton-barto:book}
in (\ref{mspbe}), an auxiliary variable $w$ is defined
\begin{equation}
w=\mathbb{E}{[\phi_t{\phi_t^{T}}]^{-1}}\mathbb{E}[\delta_{t}(\theta)\phi_t]={({\Phi^{T}}\Xi\Phi)^{-1}}{\Phi^{T}}\Xi(T\Phi\theta-\Phi\theta)
\label{eq:w}
\end{equation}
Thus we can have the following linear inverse problem
\begin{equation}
\mathbb{E}[\delta_{t}(\theta)\phi_t]=\mathbb{E}[\phi_t{\phi_t^{T}}]w=({\Phi^{T}}\Xi\Phi)w={\Phi^{T}}\Xi(T\Phi\theta-\Phi\theta)
\label{eq:tdc_lip1}
\end{equation}
By taking gradient w.r.t $\theta$ for optimum condition $\nabla{\rm {MSPBE}}(\theta)=0$
and utilizing Equation (\ref{eq:w}), we have
\begin{equation}
\mathbb{E}[\delta_{t}(\theta)\phi_t]=\gamma\mathbb{E}[\phi'_t{\phi_t^{T}}]w
\label{eq:tdc_lip2}
\end{equation}
Rearranging the two equality of Equation (\ref{eq:tdc_lip1},\ref{eq:tdc_lip2}),
we have the following linear system equation
\begin{equation}
\left[{\begin{array}{cc}
{\eta{\Phi^{T}}\Xi\Phi} & {\eta{\Phi^{T}}\Xi(\Phi-\gamma\Phi')}\\
{\gamma{\Phi^{'}}^{T}\Xi\Phi} & {{\Phi^{T}}\Xi(\Phi-\gamma\Phi')}
\end{array}}\right]\left[{\begin{array}{c}
w\\
\theta
\end{array}}\right]=\left[{\begin{array}{c}
{\eta{\Phi^{T}}\Xi R}\\
{{\Phi^{T}\Xi}R}
\end{array}}\right]\label{eq:eax=00003Db}
\end{equation}
The stochastic gradient version of the above equation is as follows,
where
\begin{equation}
A=\mathbb{E}[{A_{t}}],b=\mathbb{E}[{b_{t}}],x=[w;\theta]\label{eq:UNIFY}
\end{equation}
\begin{equation}
{A_{t}}=\left[{\begin{array}{cc}
{\eta{\phi_{t}}{\phi_{t}}^{T}} & {\eta{\phi_{t}}{{({\phi_{t}}-\gamma{{\phi'}_{t}})}^{T}}}\\
{\gamma{{\phi'}_{t}}{\phi_{t}}^{T}} & {{\phi_{t}}{{({\phi_{t}}-\gamma{{\phi'}_{t}})}^{T}}}
\end{array}}\right],{b_{t}}=\left[{\begin{array}{c}
{\eta{r_{t}}{\phi_{t}}}\\
{{r_{t}}{\phi_{t}}}
\end{array}}\right]\label{eq:abc}
\end{equation}
Following \cite{FastGradient:2009}, the TDC algorithm solution follows
from the linear equation $Ax=b$, where a single iteration gradient
update would be
\[
{x_{t+1}}={x_{t}}-{\alpha_{t}}({A_{t}}{x_{t}}-{b_{t}})
\]
where $x_{t}=[w_{t};\theta_{t}]$. The two time-scale gradient descent
learning method TDC \cite{FastGradient:2009} is
\begin{equation}
{\theta_{t+1}}={\theta_{t}}+{\alpha_{t}}{\delta_{t}}{\phi_{t}}-{\alpha_{t}}\gamma{\phi_{t}}^{\prime}(\phi_{t}^{T}{w_{t}}),{w_{t+1}}={w_{t}}+{\beta_{t}}({\delta_{t}}-\phi_{t}^{T}{w_{t}}){\phi_{t}}\label{eq:TDCupdate}
\end{equation}
where $-{\alpha_{t}}\gamma{\phi_{t}}^{\prime}(\phi_{t}^{T}{w_{t}})$
is the term for correction of gradient descent direction, and ${\beta_{t}}=\eta{\alpha_{t}},\eta>1$.
\begin{figure}[tbh]
\center
\fbox{ %
\begin{tabular}{p{11.4cm}}
\begin{itemize}
\item $\Xi$ is a diagonal matrix whose entries $\xi(s)$ are given by a
positive probability distribution over states. $\Pi=\Phi{({\Phi^{T}}\Xi\Phi)^{-1}}{\Phi^{T}}\Xi$
is the weighted least-squares projection operator.
\item A square root of $A$ is a matrix $B$ satisfying $B^{2}=A$ and $B$
is denoted as ${A^{\frac{1}{2}}}$. Note that ${A^{\frac{1}{2}}}$
may not be unique.
\item $[\cdot,\cdot]$ is a row vector, and $[\cdot;\cdot]$ is a column
vector.
\item For the $t$-th sample, $\phi_{t}$ (the $t$-th row of $\Phi$),
$\phi'_{t}$ (the $t$-th row of $\Phi'$) are the feature vectors
corresponding to $s_{t},s'_{t}$, respectively. $\theta_{t}$ is the
coefficient vector for $t$-th sample in first-order TD learning methods,
and ${\delta_{t}}=({r_{t}}+\gamma\phi_{t}^{'T}{\theta_{t}})-\phi_{t}^{T}{\theta_{t}}$
is the temporal difference error. Also, ${x_{t}}=[{w_{t}};{\theta_{t}}]$,
$\alpha_{t}$ is a stepsize, ${\beta_{t}}=\eta{\alpha_{t}},\eta>0$.
\item $m,n$ are conjugate numbers if $\frac{1}{m}+\frac{1}{n}=1,m\ge1,n\ge1$.
$||x||{_{m}}={(\sum\nolimits _{j}{|{x_{j}}{|^{m}}})^{\frac{1}{m}}}$
is the $m$-norm of vector $x$.
\item $\rho$ is $l_{1}$ regularization parameter, $\lambda$ is the eligibility
trace factor, $N$ is the sample size, $d$ is the number of basis
functions, $k$ is the number of active basis functions. \end{itemize}
\tabularnewline
\end{tabular}}
\caption{Notations and Definitions.}
\label{fig:notationmsma}
\end{figure}
There are some issues regarding the objective function, which arise
from the online convex optimization and reinforcement learning perspectives,
respectively. The first concern is that the objective function should
be convex and stochastically solvable. Note that $A,A_{t}$ are neither
PSD nor symmetric, and it is not straightforward to formulate a convex
objective function based on them. The second concern is that since
we do not have knowledge of $A$, the objective function should be
separable so that it is stochastically solvable based on $A_{t},b_{t}$.
The other concern regards the sampling condition in temporal difference
learning: double-sampling. As pointed out in \cite{sutton-barto:book},
double-sampling is a necessary condition to obtain an unbiased estimator
if the objective function is the Bellman residual or its derivatives
(such as projected Bellman residual), wherein the product of Bellman
error or projected Bellman error metrics are involved. To overcome
this sampling condition constraint, the product of TD errors should
be avoided in the computation of gradients. Consequently, based on
the linear equation formulation in (\ref{eq:UNIFY}) and the requirement
on the objective function discussed above, we propose the regularized
loss function as
\begin{equation}
L(x)={\left\Vert {Ax-b}\right\Vert _{m}}+h(x)\label{eq:lossfunc}
\end{equation}
Here we also enumerate some intuitive objective functions and give
a brief analysis on the reasons why they are not suitable for regularized
off-policy first-order TD learning. One intuitive idea is to add a
sparsity penalty on MSPBE, i.e., $L(\theta)={\rm {MSPBE(}}\theta{\rm {)+}}\rho{\left\Vert \theta\right\Vert _{1}}.$
Because of the $l_{1}$ penalty term, the solution to $\nabla L=0$
does not have an analytical form and is thus difficult to compute.
The second intuition is to use the online least squares formulation
of the linear equation $Ax=b$. However, since $A$ is not symmetric
and positive semi-definite (PSD), ${A^{\frac{1}{2}}}$ does not exist
and thus $Ax=b$ cannot be reformulated as ${\min_{x\in X}}||{{A^{\frac{1}{2}}}x-{A^{-\frac{1}{2}}}b}||_{2}^{2}$.
Another possible idea is to attempt to find an objective function
whose gradient is exactly ${A_{t}}{x_{t}}-{b_{t}}$ and thus the regularized
gradient is $pro{x_{{\alpha_{t}}h({x_{t}})}}({A_{t}}{x_{t}}-{b_{t}})$.
However, since $A_{t}$ is not symmetric, this gradient does not explicitly
correspond to any kind of optimization problem, not to mention a convex
one{%
\footnote{Note that the $A$ matrix in GTD2's linear equation representation
is symmetric, yet is not PSD, so it cannot be formulated as a convex
problem.%
}}.
\subsection{Squared Loss Formulation\label{sec:Extension}}
It is also worth noting that there exists another formulation of the
loss function different from Equation (\ref{eq:lossfunc}) with the
following convex-concave formulation as in \cite{nesterov:composite:2007,sra2011optimization},
\begin{eqnarray}
\mathop{\min}\limits _{x}\frac{1}{2}\left\Vert {Ax-b}\right\Vert _{2}^{2}+\rho{\left\Vert x\right\Vert _{1}} & = & \mathop{\max}\limits _{{{\left\Vert {{A^{T}}y}\right\Vert }_{\infty}}\le1}({b^{T}}y-\frac{\rho}{2}{y^{T}}y)\nonumber \\
& = & \mathop{\min}\limits _{x}\mathop{\max}\limits _{{{\left\Vert u\right\Vert }_{\infty}}\le1,y}\left({{x^{T}}u+{y^{T}}(Ax-b)-\frac{\rho}{2}{y^{T}}y}\right)\label{eq:formula2}
\end{eqnarray}
\label{eq:ROTD2}
Here we give the detailed deduction of formulation in Equation (\ref{eq:formula2}).
First, using the dual norm representation, the standard LASSO problem
formulation is reformulated as
\begin{equation}
f(x)=\frac{1}{2}\left\Vert {Ax-b}\right\Vert _{2}^{2}+\rho{\left\Vert x\right\Vert _{1}}=\mathop{\max}\limits _{y,{{\left\Vert {{A^{T}}y}\right\Vert }_{\infty}}\le1}\left[{\left\langle {b/\rho,y}\right\rangle -\frac{1}{2}{y^{T}}y}\right]\label{eq:formula3}
\end{equation}
Then%
\footnote{Let $w=-y$, then we will have the same formulation as in Nemirovski's
tutorial in COLT2012.
\[
\Phi(x,w)=\left\langle {w,Ax-b}\right\rangle -\frac{1}{2}{w^{T}}w-\left\langle {x,{A^{T}}w}\right\rangle
\]
}
\[
\begin{array}{l}
\left\langle {b,y}\right\rangle -\frac{1}{2}{y^{T}}y=\left\langle {b,y}\right\rangle -\frac{1}{2}{y^{T}}y+\left\langle {x,{A^{T}}y}\right\rangle -\left\langle {y,Ax}\right\rangle \\
=\left\langle {y,b-Ax}\right\rangle -\frac{1}{2}{y^{T}}y+\left\langle {x,{A^{T}}y}\right\rangle
\end{array}
\]
which can be solved iteratively without the proximal gradient step
as follows, which serves as a counterpart of Equation (\ref{eq:RO-TD}),
\begin{eqnarray}
{x_{t+1}}={x_{t}}-{\alpha_{t}}\rho({u_{t}}+{A_{t}}^{T}{y_{t}}) & , & {y_{t+1}}={y_{t}}+\frac{{\alpha_{t}}}{\rho}({A_{t}}{x_{t}}-{b_{t}}-\rho{y_{t}})\nonumber \\
{u_{t+\frac{1}{2}}}={u_{t}}+\frac{{\alpha_{t}}}{\rho}{x_{t}} & , & {u_{t+1}}={\Pi_{\infty}}({u_{t+\frac{1}{2}}})
\end{eqnarray}
\section{Algorithm Design}
\subsection{RO-TD Algorithm Design \label{sec:stochasticsaddlepoint}}
In this section, the problem of (\ref{eq:lossfunc}) is formulated
as a convex-concave saddle-point problem, and the RO-TD algorithm
is proposed. Analogous to (\ref{eq:minimax}), the regularized loss
function can be formulated as
\begin{equation}
{\left\Vert {Ax-b}\right\Vert _{m}}+h(x)={\max_{{{\left\Vert y\right\Vert }_{n}}\le1}}{y^{T}}(Ax-b)+h(x)\label{eq:regminimax}
\end{equation}
Similar to (\ref{eq:minimaxFOM}), Equation (\ref{eq:regminimax})
can be solved via an iteration procedure as follows, where ${{x}_{t}}=\left[{{w_{t}};{\theta_{t}}}\right]$.
\begin{eqnarray}
{{x}_{t+\frac{1}{2}}}={x_{t}}-{\alpha_{t}}{A_{t}^{T}}{y_{t}} & , & {{y}_{t+\frac{1}{2}}}={y_{t}}+\alpha_{t}({A_{t}}{x_{t}}-{b_{t}})\nonumber \\
{x_{t+1}}=pro{x_{{\alpha_{t}}h}}({x_{t+\frac{1}{2}}}) & , & {y_{t+1}}={\Pi_{n}}({y_{t+\frac{1}{2}}})\label{eq:RO-TD}
\end{eqnarray}
The averaging step, which plays a crucial role in stochastic optimization
convergence, generates the \emph{approximate saddle-points} \cite{sra2011optimization,nedic2009subgradient}
\begin{equation}
{\bar{x}_{t}}={\left({\sum\nolimits _{i=0}^{t}{\alpha_{i}}}\right)^{-1}}\sum\nolimits _{i=0}^{t}{{\alpha_{i}}{x_{i}}},{\bar{y}_{t}}={\left({\sum\nolimits _{i=0}^{t}{\alpha_{i}}}\right)^{-1}}\sum\nolimits _{i=0}^{t}{{\alpha_{i}}{y_{i}}}\label{eq:averaging}
\end{equation}
Due to the computation of $A_{t}$ in (\ref{eq:RO-TD}) at each iteration,
the computation cost appears to be $O(Nd^{2})$, where $N,d$ are
defined in Figure \ref{fig:notationmsma}. However, the computation
cost is actually $O(Nd)$ with a linear algebraic trick by computing
not $A_{t}$ but $y_{t}^{T}{A_{t}},{A_{t}}{x_{t}}-{b_{t}}$. Denoting
${y_{t}}=[{y_{1,t}};{y_{2,t}}]$, where ${y_{1,t}};{y_{2,t}}$ are
column vectors of equal length, we have
\begin{equation}
y_{t}^{T}{A_{t}}=\left[{\begin{array}{cc}
{\eta\phi_{t}^{T}(y_{1,t}^{T}{\phi_{t}})+\gamma\phi_{t}^{T}(y_{2,t}^{T}\phi_{t}^{\prime})} & {{{({\phi_{t}}-\gamma\phi_{t}^{\prime})}^{T}}(\eta y_{1,t}^{T}+y_{2,t}^{T}){\phi_{t}}}\end{array}}\right]\label{eq:ytAt}
\end{equation}
${A_{t}}{x_{t}}-{b_{t}}$ can be computed according to Equation (\ref{eq:TDCupdate})
as follows:
\begin{equation}
{A_{t}}{x_{t}}-{b_{t}}=\left[{\begin{array}{c}
{-\eta({\delta_{t}}-\phi_{t}^{T}{w_{t}}){\phi_{t}}};{\gamma(\phi_{t}^{T}{w_{t}}){\phi_{t}}^{\prime}-{\delta_{t}}{\phi_{t}}}\end{array}}\right]\label{eq:Atxt-bt}
\end{equation}
Both (\ref{eq:ytAt}) and (\ref{eq:Atxt-bt}) are of linear computational
complexity. Now we are ready to present the RO-TD algorithm:
There are some design details of the algorithm to be elaborated. First,
the regularization term $h(x)$ can be any kind of convex regularization,
such as ridge regression or sparsity penalty $\rho||x||_{1}$. In
case of $h(x)=\rho||x||_{1}$, $pro{x_{{\alpha_{t}}h}}(\cdot)={S_{{\alpha_{t}}\rho}}(\cdot)$.
In real applications the sparsification requirement on $\theta$ and
auxiliary variable $w$ may be different, i.e., $h(x)={\rho_{1}}{\left\Vert \theta\right\Vert _{1}}+{\rho_{2}}{\left\Vert w\right\Vert _{1}},{\rho_{1}}\ne{\rho_{2}}$,
one can simply replace the uniform soft thresholding $S_{{\alpha_{t}}\rho}$
by two separate soft thresholding operations $S_{{\alpha_{t}}\rho_{1}},S_{{\alpha_{t}}\rho_{2}}$
and thus the third equation in (\ref{eq:RO-TD}) is replaced by the
following,
\begin{equation}
{x_{t+\frac{1}{2}}}=\left[{{w_{t+\frac{1}{2}}};{\theta_{t+\frac{1}{2}}}}\right],{\theta_{t+1}}={S_{{\alpha_{t}}{\rho_{1}}}}({\theta_{t+\frac{1}{2}}}),{w_{t+1}}={S_{{\alpha_{t}}{\rho_{2}}}}({w_{t+\frac{1}{2}}})\label{twothreshold}
\end{equation}
Another concern is the choice of conjugate numbers $(m,n)$. For ease
of computing $\Pi_{n}$, we use $(2,2)$($l_{2}$ fit), $(+\infty,1)$(uniform
fit) or $(1,+\infty)$. $m=n=2$ is used in the experiments below.
\begin{algorithm}[tph]
\caption{RO-TD}
\label{alg: RO-TD}
Let $\pi$ be some fixed policy of an MDP $M$, and let the sample
set $S=\{{s_{i}},{r_{i}},{s_{i}}'\}_{i=1}^{N}$. Let $\Phi$ be some
fixed basis.
\begin{enumerate}
\item \textbf{REPEAT}
\item Compute ${{\phi_{t}},{\phi_{t}}'}$ and TD error ${\delta_{t}}=({r_{t}}+\gamma\phi_{t}^{'T}{\theta_{t}})-\phi_{t}^{T}{\theta_{t}}$
\item Compute $y_{_{t}}^{T}{A_{t}},{A_{t}}{x_{t}}-{b_{t}}$ in Equation
(\ref{eq:ytAt}) and (\ref{eq:Atxt-bt}).
\item Compute $x_{t+1},y_{t+1}$ as in Equation (\ref{eq:RO-TD})
\item Set $t\leftarrow t+1$;
\item \textbf{UNTIL} {$t=N$};
\item Compute $\bar{x}_{N},\bar{y}_{N}$ as in Equation (\ref{eq:averaging})
with $t=N$ .\end{enumerate}
\end{algorithm}
\subsection{RO-GQ($\lambda$) Design}
GQ($\lambda$)\cite{gq:maei2010} is a generalization of the TDC algorithm
with eligibility traces and off-policy learning of temporally abstract
predictions, where the gradient update changes from Equation (\ref{eq:TDCupdate})
to
\begin{equation}
{\theta_{t+1}}={\theta_{t}}+{\alpha_{t}}[{\delta_{t}}{e_{t}}-\gamma(1-\lambda){w_{t}}^{T}{e_{t}}{\bar{\phi}_{t+1}}],{w_{t+1}}={w_{t}}+{\beta_{t}}({\delta_{t}}{e_{t}}-w_{t}^{T}{\phi_{t}}{\phi_{t}})\label{eq:gq}
\end{equation}
The central element is to extend the MSPBE function to the case where
it incorporates eligibility traces. The objective function and corresponding
linear equation component $A_{t},b_{t}$ can be written as follows:
\begin{equation}
L(\theta)=||\Phi\theta-\Pi T^{\pi\lambda}\Phi\theta||_{\Xi}^{2}\label{eq:gqObj}
\end{equation}
\begin{equation}
{A_{t}}=\left[{\begin{array}{cc}
{\eta{\phi_{t}}{\phi_{t}}^{T}} & {\eta{e_{t}}{{({\phi_{t}}-\gamma{\bar{\phi}_{t+1}})}^{T}}}\\
{\gamma(1-\lambda){{\bar{\phi}}_{t+1}}e_{t}^{T}} & {{e_{t}}{{({\phi_{t}}-\gamma{\bar{\phi}_{t+1}})}^{T}}}
\end{array}}\right],{b_{t}}=\left[{\begin{array}{c}
{\eta{r_{t}}{e_{t}}}\\
{{r_{t}}{e_{t}}}
\end{array}}\right]\label{eq:gqAb}
\end{equation}
Similar to Equation (\ref{eq:ytAt}) and (\ref{eq:Atxt-bt}), the
computation of $y_{_{t}}^{T}{A_{t}},{A_{t}}{x_{t}}-{b_{t}}$ is
\begin{eqnarray}
y_{_{t}}^{T}{A_{t}} & = & \left[{\begin{array}{cc}
{\eta\phi_{t}^{T}(y_{1,t}^{T}{\phi_{t}})+\gamma(1-\lambda)e_{t}^{T}(y_{2,t}^{T}{\bar{\phi}_{t+1}})} & {{({\phi_{t}}-\gamma{\bar{\phi}_{t+1}})^{T}}(\eta y_{1,t}^{T}+y_{2,t}^{T}){e_{t}}}\end{array}}\right]\nonumber \\
{A_{t}}{x_{t}}-{b_{t}} & = & \left[{\begin{array}{c}
{-\eta({\delta_{t}}{e_{t}}-\phi_{t}^{T}{w_{t}}{\phi_{t}})};{\gamma(1-\lambda)(e_{t}^{T}{w_{t}}){\bar{\phi}_{t+1}}-{\delta_{t}}{e_{t}}}\end{array}}\right]\label{eq:gqcase}
\end{eqnarray}
where eligibility traces $e_{t}$, and ${\bar{\phi}_{t}},T^{\pi\lambda}$
are defined in \cite{gq:maei2010}. Algorithm \ref{alg: RO-GQ}, RO-GQ($\lambda$),
extends the RO-TD algorithm to include eligibility traces.
\begin{algorithm}[tph]
\caption{RO-GQ($\lambda$)}
\label{alg: RO-GQ}
Let $\pi$ be some fixed policy of an MDP $M$. Let $\Phi$ be some
fixed basis. Starting from $s_{0}$.
\begin{enumerate}
\item \textbf{REPEAT}
\item Compute ${{\phi_{t}},{\bar{\phi}_{t+1}}}$ and TD error ${\delta_{t}}=({r_{t}}+\gamma\bar{\phi}_{t+1}^{T}{\theta_{t}})-\phi_{t}^{T}{\theta_{t}}$
\item Compute $y_{_{t}}^{T}{A_{t}},{A_{t}}{x_{t}}-{b_{t}}$ in Equation
(\ref{eq:gqcase}).
\item Compute $x_{t+1},y_{t+1}$ as in Equation (\ref{eq:RO-TD})
\item Choose action $a_{t}$, and get $s_{t+1}$
\item Set $t\leftarrow t+1$;
\item \textbf{UNTIL} $s_{t}$ is an absorbing state;
\item Compute $\bar{x}_{t},\bar{y}_{t}$ as in Equation (\ref{eq:averaging}) \end{enumerate}
\end{algorithm}
\section{Theoretical Analysis}
The theoretical analysis of RO-TD algorithm can be seen in the Appendix.
\section{Empirical Results\label{sec:Empirical-Experiments-1}}
We now demonstrate the effectiveness of the RO-TD algorithm against
other algorithms across a number of benchmark domains. LARS-TD \cite{Kolter09LARSTD},
which is a popular second-order sparse reinforcement learning algorithm,
is used as the baseline algorithm for feature selection and TDC is
used as the off-policy convergent RL baseline algorithm, respectively.
\subsection{MSPBE Minimization and Off-Policy Convergence}
\begin{figure}[tph]
\centering %
\begin{minipage}[c]{1\textwidth}%
\begin{center}
\includegraphics[width=4in,height=2in]{Star1}
\par\end{center}
\begin{center}
\includegraphics[width=4in,height=2in]{Star2}
\par\end{center}
\begin{center}
\includegraphics[width=4in,height=2in]{randomwalk}
\caption{Illustrative examples of the convergence of RO-TD using the Star and
Random-walk MDPs.}
\label{fig:STAR}
\par\end{center}
\end{minipage}
\end{figure}
This experiment aims to show the minimization of MSPBE and off-policy
convergence of the RO-TD algorithm. The $7$ state star MDP is a well
known counterexample where TD diverges monotonically and TDC converges.
It consists of $7$ states and the reward w.r.t any transition is
zero. Because of this, the star MDP is unsuitable for LSTD-based algorithms,
including LARS-TD since ${\Phi^{T}}R=0$ always holds. The random-walk
problem is a standard Markov chain with $5$ states and two absorbing
state at two ends. Three sets of different bases $\Phi$ are used
in \cite{FastGradient:2009}, which are tabular features, inverted
features and dependent features respectively. An identical experiment
setting to \cite{FastGradient:2009} is used for these two domains.
The regularization term $h(x)$ is set to $0$ to make a fair comparison
with TD and TDC. $\alpha=0.01$, $\eta=10$ for TD, TDC and RO-TD.
The comparison with TD, TDC and RO-TD is shown in the left sub-figure
of Figure \ref{fig:STAR}, where TDC and RO-TD have almost identical
MSPBE over iterations. The middle sub-figure shows the value of $y_{_{t}}^{T}(A{x_{t}}-b)$
and ${\left\Vert {A{x_{t}}-b}\right\Vert _{2}}$, wherein ${\left\Vert {A{x_{t}}-b}\right\Vert _{2}}$
is always greater than the value of $y_{_{t}}^{T}(A{x_{t}}-b)$. Note
that for this problem, the Slater condition is satisfied so there
is no duality gap between the two curves. As the result shows, TDC
and RO-TD perform equally well, which illustrates the off-policy convergence
of the RO-TD algorithm. The result of random-walk chain is averaged
over $50$ runs. The rightmost sub-figure of Figure \ref{fig:STAR}
shows that RO-TD is able to reduce MSPBE over successive iterations
w.r.t three different basis functions.
\subsection{Feature Selection}
In this section, we use the mountain car example with a variety of
bases to show the feature selection capability of RO-TD. The Mountain
car is an optimal control problem with a continuous two-dimensional
state space. The steep discontinuity in the value function makes learning
difficult for bases with global support. To make a fair comparison,
we use the same basis function setting as in \cite{Kolter09LARSTD},
where two dimensional grids of $2,4,8,16,32$ RBFs are used so that
there are totally $1365$ basis functions. For LARS-TD, $500$ samples
are used. For RO-TD and TDC, $3000$ samples are used by executing
$15$ episodes with $200$ steps for each episode, stepsize $\alpha_{t}=0.001$,
and $\rho_{1}=0.01,\rho_{2}=0.2$. We use the result of LARS-TD and
$l_{2}$ LSTD reported in \cite{Kolter09LARSTD}. As the result shows
in Table~\ref{tab:mcar}, RO-TD is able to perform feature selection
successfully, whereas TDC and TD failed. It is worth noting that comparing
the performance of RO-TD and LARS-TD is not the major focus here,
since LARS-TD is not convergent off-policy and RO-TD's performance
can be further optimized using the mirror-descent approach with the
Mirror-Prox algorithm \cite{sra2011optimization} which incorporates
mirror descent with an extragradient \cite{extragradient:1976}, as
discussed below.
\begin{table}[h]
\begin{centering}
\centering %
\begin{tabular}{|c|c|c|c|c|c|}
\hline
Algorithm & LARS-TD & RO-TD & $l_{2}$ LSTD & TDC & TD\tabularnewline
\hline
Success($20/20$) & $100\%$ & $100\%$ & $0\%$ & $0\%$ & $0\%$ \tabularnewline
\hline
Steps & $142.25\pm9.74$ & $147.40\pm13.31$ & - & - & - \tabularnewline
\hline
\end{tabular}\caption{Comparison of TD, LARS-TD, RO-TD, $l_{2}$ LSTD, TDC and TD}
\label{tab:mcar}
\par\end{centering}
\end{table}
\begin{table}[tph]
\begin{centering}
\centering %
\begin{tabular}{|c|c|c|c|}
\hline
Experiment\textbackslash{}Method & RO-GQ($\lambda$) & GQ($\lambda$) & LARS-TD \tabularnewline
\hline
Experiment 1 & $6.9\pm4.82$ & $11.3\pm9.58$ & - \tabularnewline
\hline
Experiment 2 & $14.7\pm10.70$ & $27.2\pm6.52$ & - \tabularnewline
\hline
\end{tabular}\caption{Comparison of RO-GQ($\lambda$), GQ($\lambda$), and LARS-TD on Triple-Link
Inverted Pendulum Task
\label{tab:rogq-triple}
\par\end{centering}
\end{table}
\subsection{High-dimensional Under-actuated Systems}
The triple-link inverted pendulum \cite{si2001online} is a highly
nonlinear under-actuated system with $8$-dimensional state space
and discrete action space. The state space consists of the angles
and angular velocity of each arm as well as the position and velocity
of the car. The discrete action space is $\{0,5{\rm {Newton}},-5{\rm {Newton}}\}$.
The goal is to learn a policy that can balance the arms for $N_{x}$
steps within some minimum number of learning episodes. The allowed
maximum number of episodes is $300$. The pendulum initiates from
zero equilibrium state and the first action is randomly chosen to
push the pendulum away from initial state. We test the performance
of RO-GQ($\lambda$), GQ($\lambda$) and LARS-TD. Two experiments
are conducted with $N_{x}=10,000$ and $100,000$, respectively. Fourier
basis \cite{konidaris:fourier} with order $2$ is used, resulting
in $6561$ basis functions. Table \ref{tab:rogq-triple} shows the results of this experiment,
where RO-GQ($\lambda$) performs better than other approaches, especially
in Experiment 2, which is a harder task. LARS-TD failed in this domain,
which is mainly not due to LARS-TD itself but the quality of samples
collected via random walk.
To sum up, RO-GQ($\lambda$) tends to outperform GQ($\lambda$) in
all aspects, and is able to outperform LARS-TD based policy iteration
in high dimensional domains, as well as in selected smaller MDPs where
LARS-TD diverges (e.g., the star MDP). It is worth noting that the
computation cost of LARS-TD is $O(Ndk^{2})$, where that for RO-TD
is $O(Nd)$. If $k$ is linear or sublinear w.r.t $d$, RO-TD has
a significant advantage over LARS-TD. However, compared with LARS-TD,
RO-TD requires fine tuning the parameters of $\alpha_{t},\rho_{1},\rho_{2}$
and is usually not as sample efficient as LARS-TD. We also find that
tuning the sparsity parameter $\rho_{2}$ generates an interpolation
between GQ($\lambda$) and Q-learning, where a large $\rho_{2}$
helps eliminate the correction term of TDC update and make the update
direction more similar to the TD update.
\section{Summary}
In this chapter we present a novel unified framework for designing regularized
off-policy convergent RL algorithms combining a convex-concave saddle-point
problem formulation for RL with stochastic first-order methods. A
detailed experimental analysis reveals that the proposed RO-TD algorithm
is both off-policy convergent and robust to noisy features.
\chapter{Safe Reinforcement Learning using Projected Natural Actor Critic}
\label{pnac-chapter}
Natural actor-critics form a popular class of policy search algorithms for finding locally optimal policies for Markov decision processes. In this paper we address a drawback of natural actor-critics that limits their real-world applicability---their lack of safety guarantees. We present a principled algorithm for performing natural gradient descent over a constrained domain \footnote{This paper is a revised version of the paper ``Projected Natural Actor-Critic" that was published in NIPS 2013.}. In the context of reinforcement learning, this allows for natural actor-critic algorithms that are guaranteed to remain within a known safe region of policy space. While deriving our class of constrained natural actor-critic algorithms, which we call Projected Natural Actor-Critics (PNACs), we also elucidate the relationship between natural gradient descent and mirror descent.
\section{Introduction}
Natural actor-critics form a class of policy search algorithms for finding locally optimal policies for Markov decision processes (MDP
) by approximating and ascending the natural gradient \cite{Amari1998} of an objective function. Despite the numerous successes of, and the continually growing interest in, natural actor-critic algorithms, they have not achieved widespread use for real-world applications. A lack of safety guarantees is a common reason for avoiding the use of natural actor-critic algorithms, particularly for biomedical applications. Since natural actor-critics are \emph{unconstrained} optimization algorithms, there are no guarantees that they will avoid regions of policy space that are known to be dangerous.
For example, proportional-integral-derivative controllers (PID controllers) are the most widely used control algorithms in industry, and have been studied in depth \cite{Astrom1995}.
Techniques exist for determining the set of stable gains (policy parameters) when a model of the system is available \cite{Soylemez2003}. Policy search can be used to find the optimal gains within this set (for some definition of optimality).
A desirable property of a policy search algorithm in this context would be a guarantee that it will remain within the predicted region of stable gains during its search.
Consider a second example: \emph{functional electrical stimulation} (FES) control of a human arm. By selectively stimulating muscles using subcutaneous probes, researchers have made significant strides toward returning motor control to people suffering from paralysis induced by spinal cord injury \cite{Lynch2008}. There has been a recent push to develop controllers that specify how much and when to stimulate each muscle in a human arm to move it from its current position to a desired position \cite{Chadwick2009}. This closed-loop control problem is particularly challenging because each person's arm has different dynamics due to differences in, for example, length, mass, strength, clothing, and amounts of muscle atrophy, spasticity, and fatigue. Moreover, these differences are challenging to model. Hence, a proportional-derivative (PD) controller, tuned to a simulation of an ideal human arm, required manual tuning to obtain desirable performance on a human subject with biceps spasticity \cite{KathyPD}.
Researchers have shown that policy search algorithms are a viable approach to creating controllers that can automatically adapt to an individual's arm by training on a few hundred two-second reaching movements \cite{ThomasIAAI}.
However, safety concerns have been raised in regard to both this specific application and other biomedical applications of policy search algorithms. Specifically, the existing state-of-the-art gradient-based algorithms, including the current natural actor-critic algorithms, are unconstrained and could potentially select dangerous policies. For example, it is known that certain muscle stimulations could cause the dislocation of a subject's arm. Although we lack an accurate model of each individual's arm, we can generate conservative safety constraints on the space of policies. Once again, a desirable property of a policy search algorithm would be a guarantee that it will remain within a specified region of policy space (known-safe policies).
In this paper we present a class of natural actor-critic algorithms that perform constrained optimization---given a known safe region of policy space, they search for a locally optimal policy while always remaining within the specified region. We call our class of algorithms \emph{Projected Natural Actor-Critics} (PNACs) since, whenever they generate a new policy, they project the policy back to the set of safe policies. The interesting question is how the projection can be done in a principled manner. We show that natural gradient descent (ascent), which is an \emph{unconstrained} optimization algorithm, is a special case of mirror descent (ascent), which is a \emph{constrained} optimization algorithm. In order to create a projected natural gradient algorithm, we add constraints in the mirror descent algorithm that is equivalent to natural gradient descent. We apply this projected natural gradient algorithm to policy search to create the PNAC algorithms, which we validate empirically.
\section{Related Work}
Researchers have addressed safety concerns like these before \cite{Perkins200
}. \citet{Bendrahim1997} showed how a walking biped robot can switch to a stabilizing controller whenever the robot leaves a stable region of state space. Similar state-avoidant approaches to safety have been proposed by several others \cite{Arapostathis2005
Arvelo2012
Geibel2005}. These approaches do not account for situations where, over an unavoidable region of state space, the actions themselves are dangerous. \citet{Kuindersma2012} developed a method for performing risk-sensitive policy search, which models the variance of the objective function for each policy and permits runtime adjustments of risk sensitivity. However, their approach does not guarantee that an unsafe region of state space or policy space will be avoided.
\citet{Bhatnagar2009} presented projected natural actor-critic algorithms for the average reward setting. As in our projected natural actor-critic algorithms, they proposed computing the update to the policy parameters and then projecting back to the set of allowed policy parameters. However, they did not specify how the projection could be done in a principled manner. We show in Section \ref{sec:CP} that the Euclidean projection can be arbitrarily bad, and argue that the projection that we propose is particularly compatible with natural actor-critics (natural gradient descent).
\citet{Duchi2010} presented mirror descent using the Mahalanobis norm for the proximal function, which is very similar to the proximal function that we show to cause mirror descent to be equivalent to natural gradient descent. However, their proximal function is not identical to ours and they did not discuss any possible relationship between mirror descent and natural gradient descent.
\section{Equivalence of Natural Gradient Descent and Mirror Descent}
We begin by showing an important relationship between natural gradient methods and mirror descent.
\begin{thm}
The natural gradient descent update at step $k$ with metric tensor $G_k \triangleq G(x_k)$:
\begin{equation}
x_{k+1} = x_k - \alpha_k G_k^{-1}\nabla f(x_k),
\label{eq:NG}
\end{equation}
is equivalent to the mirror descent update at step $k$, with $\psi_k(x)=(\sfrac{1}{2})x^\intercal G_k x$.
\end{thm}
\begin{proof}
First, notice that $\nabla \psi_k(x) = G_kx$. Next, we derive a closed-form for $\psi_k^*$:
\begin{align}
\psi_k^*(y) = \max_{x \in \mathbb R^n} \left \{ x^\intercal y - \frac{1}{2}x^\intercal G_k x \right \}.
\label{eq:proof1}
\end{align}
Since the function being maximized on the right hand side is strictly concave, the $x$ that maximizes it is its critical point. Solving for this critical point, we get $x = G_k^{-1}y$. Substituting this into \eqref{eq:proof1}, we find that
$
\psi_k^*(y)=(\sfrac{1}{2})y^\intercal G_k^{-1}y.
$
Hence, $\nabla \psi_k^*(y) = G_k^{-1}y$. Using the definitions of $\nabla \psi_k(x)$ and $\nabla \psi_k^*(y)$, we find that the mirror descent update is
\begin{align}
x_{k+1}=& G_k^{-1} \left ( G_k x_k - \alpha_k \nabla f(x_k) \right )= x_k -\alpha_k G_k^{-1}\nabla f(x_k),
\end{align}
which is identical to \eqref{eq:NG}.{\hfill$\blacksquare$}\end{proof}
Although researchers often use $\psi_k$ that are norms like the $p$-norm and Mahalanobis norm, notice that the $\psi_k$ that results in natural gradient descent is \emph{not} a norm. Also, since $G_k$ depends on $k$, $\psi_k$ is an \emph{adaptive} proximal function \cite{Duchi2010}.
\section{Projected Natural Gradients}
When $x$ is constrained to some set, $X$, $\psi_k$ in mirror descent is augmented with the indicator function $I_X$, where $I_X(x)=0$ if $x \in X$, and $+\infty$ otherwise. The $\psi_k$ that was shown to generate an update equivalent to the natural gradient descent update, with the added constraint that $x \in X$, is $\psi_k(x) = (\sfrac{1}{2})x^\intercal G_kx + I_X(x).$ Hereafter, any references to $\psi_k$ refer to this augmented version.
For this proximal function, the subdifferential of $\psi_k(x)$ is
$
\nabla \psi_k(x)=G_k(x) + \hat N_X(x)= (G_k+\hat N_X)(x),
$
where $\hat N_X(x) \triangleq \partial I_X(x)$ and, in the middle term, $G_k$ and $\hat N_X$ are relations and $+$ denotes Minkowski addition.\footnote{Later, we abuse notation and switch freely between treating $G_k$ as a matrix and a relation. When it is a matrix, $G_kx$ denotes matrix-vector multiplication that produces a vector. When it is a relation, $G_k(x)$ produces the singleton $\{G_kx\}$.} $\hat N_X(x)$ is the normal cone of $X$ at $x$ if $x \in X$ and $\emptyset$ otherwise \cite{Rockafellar1970}.
\begin{align}
\nabla \psi_k^*(y) = (G_k+ \hat N_X)^{-1}(y).
\label{eq:lkjhasdkltjt}
\end{align}
Let $\Pi_X^{G_k}(y)$, be the set of $x \in X$ that are closest to $y$, where the length of a vector, $z$, is $(\sfrac{1}{2})z^\intercal G_kz$. More formally,
\begin{align}
\Pi_X^{G_k}(y) \triangleq \arg \min_{x \in X} \frac{1}{2}(y-x)^\intercal G_k (y-x).
\label{eq:blah}
\end{align}
\begin{lemma}
\label{lemma:thingy}
$\Pi_X^{G_k}(y) = (G_k+\hat N_X)^{-1}(G_ky)$.
\end{lemma}
\begin{proof}
We write \eqref{eq:blah} without the explicit constraint that $x \in X$ by appending the indicator function:
\begin{align}
\Pi_X^{G_k}(y
=& \arg \min_{x \in \mathbb R^n} h_y(x),
\end{align}
where $h_y(x)=(\sfrac{1}{2}) (y-x)^\intercal G_k(y-x) + I_X(x)$. Since $h_y$ is strictly convex over $X$ and $+\infty$ elsewhere, its critical point is its global minimizer. The critical point satisfies
\begin{align}
0 \in \nabla h_y(x) =-G_k(y)+G_k(x)+\hat N_X(x).
\end{align}
The globally minimizing $x$ therefore satisfies
$
G_ky \in G_k(x) + \hat N_X(x)= (G_k+\hat N_X)(x).
$
Solving for $x$, we find that $x = (G_k+\hat N_X)^{-1}(G_ky)$.{\hfill$\blacksquare$}\end{proof}
Combining Lemma \ref{lemma:thingy}
with \eqref{eq:lkjhasdkltjt}, we find that
$
\nabla \psi^*(y)=\Pi_X^{G_k}(G_k^{-1}y).
$
Hence, mirror descent with the proximal function that produces natural gradient descent, augmented to include the constraint that $x \in X$, is:
\begin{align}
x_{k+1} =& \Pi_X^{G_k} \left (G_k^{-1} \left ( (G_k+\hat N_X)(x_k) - \alpha_k \nabla f(x_k) \right ) \right )\\
=& \Pi_X^{G_k} \left ( (I+G_k^{-1}\hat N_X)(x_k) - \alpha_k G_k^{-1} \nabla f(x_k) \right ),
\end{align}
where $I$ denotes the identity relation. Since $x_k \in X$, we know that $0 \in \hat N_X(x_k)$, and hence the update can be written as
\begin{align}
x_{k+1} = \Pi_X^{G_k} \left ( x_k - \alpha_k G_k^{-1} \nabla f(x_k) \right ),
\label{eq:PNG}
\end{align}
which we call \emph{projected natural gradient} (PNG).
\section{Compatibility of Projection}
\label{sec:CP}
The standard projected subgradient (PSG) descent method follows the negative gradient (as opposed to the negative natural gradient) and projects back to $X$ using the Euclidean norm. If $f$ and $X$ are convex and the stepsize is decayed appropriately, it is guaranteed to converge to a global minimum, $x^* \in X$. Any such $x^*$ is a fixed point. This means that a small step in the negative direction of any subdifferential of $f$ at $x^*$ will project back to $x^*$.
Our choice of projection, $\Pi_X^{G_k}$, results in PNG having the same fixed points (see Lemma \ref{lemma:compatibleProjection}). This means that, when the algorithm is at $x^*$ and a small step is taken down the natural gradient to $x'$, $\Pi_X^{G_k}$ will project $x'$ back to $x^*$. We therefore say that $\Pi_X^{G_k}$ is compatible with the natural gradient. For comparison, the Euclidean projection of $x'$ will \emph{not} necessarily return $x'$ to $x^*$.
\begin{lemma}
\label{lemma:compatibleProjection}
The sets of fixed points for PSG and PNG are equivalent.
\end{lemma}
\begin{proof}
A necessary and sufficient condition for $x$ to be a fixed point of PSG is that $-\nabla f(x) \in \hat N_X(x)$ \cite{Nocedal2006}.
A necessary and sufficient condition for $x$ to be a fixed point of PNG is
\begin{align}
x =& \Pi_X^{G_k} \left ( x - \alpha_k G_k^{-1} \nabla f(x) \right )= (G_k+\hat N_X)^{-1} \Big ( G_k \left ( x - \alpha_k G_k^{-1} \nabla f(x) \right ) \Big )\\
=& (G_k+\hat N_X)^{-1} \left ( G_kx - \alpha_k \nabla f(x) \right )\\
\Leftrightarrow &G_kx - \alpha_k \nabla f(x) \in G_k(x) + \hat N_X(x) \\
\Leftrightarrow& -\nabla f(x) \in \hat N_X(x).&\mbox{\hfill$\blacksquare$}
\end{align}
\end{proof}
To emphasize the importance of using a compatible projection, consider the following simple example. Minimize the function $f(x) = x^\intercal A x + b^\intercal x$, where $A = \mbox{diag}(1,0.01)$ and $b=[-0.2,-0.1]^\intercal$, subject to the constraints $\lVert x \rVert_1 \leq 1$ and $x \geq 0$. We implemented three algorithms, and ran each for $1000$ iterations using a fixed stepsize:
\begin{enumerate}
\vspace{-.2cm}\item {\bf PSG} - projected subgradient descent using the Euclidean projection.
\vspace{-.1cm}\item {\bf PNG} - projected natural gradient descent using $\Pi_X^{G_k}$.
\vspace{-.1cm}\item {\bf PNG-Euclid} - projected natural gradient descent using the Euclidean projection.
\end{enumerate}
\vspace{-.2cm}The results are shown in Figure
. Notice that PNG and PSG converge to the optimal solution, $x^*$. From this point, they both step in different directions, but project back to $x^*$. However, PNG-Euclid converges to a suboptimal solution (outside the domain of the figure). If $X$ were a line segment between the point that PNG-Euclid and PNG converge to, then PNG-Euclid would converge to the pessimal solution within $X$, while PSG and PNG would converge to the optimal solution within $X$. Also, notice that the natural gradient corrects for the curvature of the function and heads directly towards the global unconstrained minimum. Since the natural methods in this example use metric tensor $G=A$, which is the Hessian of $f$, they are essentially an incremental form of Newton's method. In practice, the Hessian is usually not known, and an estimate thereof is used.
\begin{SCfigure}
\centering
\includegraphics[width=0.4\columnwidth, height=0.15\textheight]{example.png}
\label{fig:example}
\caption{The thick diagonal line shows one constraint and dotted lines show projections. Solid arrows show the directions of the natural gradient and gradient at the optimal solution, $x^*$. The dashed blue arrows show PNG-Euclid's projections, and emphasize the the projections cause PNG-Euclid to move \emph{away} from the optimal solution. \vspace{0.15cm}}
\vspace{-.3cm}
\end{SCfigure}
\section{Natural Actor-Critic Algorithms}
\label{sec:NACA}
An MDP is a tuple $M=(\mathcal S, \mathcal A, \mathcal P, \mathcal R, d_0, \gamma)$, where $\mathcal S$ is a set of states, $\mathcal A$ is a set of actions, $\mathcal P(s'|s,a)$ gives the probability density of the system entering state $s'$ when action $a$ is taken in state $s$, $R(s,a)$ is the expected reward, $r$, when action $a$ is taken in state $s$, $d_0$ is the initial state distribution, and $\gamma \in [0,1)$ is a reward discount parameter. A parameterized policy, $\pi$, is a conditional probability density function---$\pi(a|s,\theta)$ is the probability density of action $a$ in state $s$ given a vector of policy parameters, $\theta \in \mathbb R^n$.
Let $J(\theta) = \mbox{E} \left [ \sum_{t=0}^\infty \gamma^t r_t | \theta \right ]$ be the \emph{discounted-reward objective} or the \emph{average reward objective function} with $J(\theta) = \lim_{n\to\infty} \frac{1}{n}\mbox{E} \left [ \sum_{t=0}^n r_t | \theta \right ]$. Given an MDP, $M$, and a parameterized policy, $\pi$, the goal is to find policy parameters that maximize one of these objectives. When the action set is continuous, the search for globally optimal policy parameters becomes intractable, so policy search algorithms typically search for locally optimal policy parameters.
Natural actor-critics, first proposed by \citet{Kakade2002}, are algorithms that estimate and ascend the natural gradient of $J(\theta)$, using the average Fisher information matrix as the metric tensor:
\vspace{-.1cm}\begin{equation}
G_k=G(\theta_k)=\mbox{E}_{s \sim d^\pi,a \sim \pi} \left [ \left (\frac{\partial}{\partial \theta_k} \log \pi(a|s,\theta_k)\right )\left(\frac{\partial}{\partial \theta_k} \log \pi(a|s,\theta_k)\right ) ^ \intercal \right ],
\end{equation}
where $d^\pi$ is a policy and objective function-dependent distribution over the state set \cite{Sutton2000}.
There are many natural actor-critics, including Natural policy gradient utilizing the Temporal Differences (NTD) algorithm \cite{Morimura2005}, Natural Actor-Critic using LSTD-Q$(\lambda)$ (NAC-LSTD) \cite{Peters2008}, Episodic Natural Actor-Critic (eNAC) \cite{Peters2008}, Natural Actor-Critic using Sarsa$(\lambda)$ (NAC-Sarsa) \cite{Thomas2012}, Incremental Natural Actor-Critic (INAC) \cite{Degris2012}, and Natural-Gradient Actor-Critic with Advantage Parameters (NGAC) \cite{Bhatnagar2009}. All of them form an estimate, typically denoted $w_k$, of the natural gradient of $J(\theta_k)$. That is, $w_k \approx G(\theta_k)^{-1}\nabla J(\theta_k)$. They then perform the policy parameter update, $\theta_{k+1} = \theta_k + \alpha_k w_k.$
\section{Projected Natural Actor-Critics}
If we are given a closed convex set, $\Theta \subseteq \mathbb R^n$, of admissible policy parameters (e.g., the stable region of gains for a PID controller), we may wish to ensure that the policy parameters remain within $\Theta$. The natural actor-critic algorithms described in the previous section do not provide such a guarantee. However, their policy parameter update equations, which are natural gradient ascent updates, can easily be modified to the projected natural gradient ascent update in \eqref{eq:PNG} by projecting the parameters back onto $\Theta$ using $\Pi_\Theta^{G(\theta_k)}$:
\begin{align}
\theta_{k+1} = \Pi_\Theta^{G(\theta_k)} \!\! \left ( \theta_k + \alpha_k w_k \right ).
\end{align}
Many of the existing natural policy gradient algorithms, including NAC-LSTD, eNAC, NAC-Sarsa, and INAC, follow \emph{biased} estimates of the natural policy gradient~\cite{Thomas2012b}. For our experiments, we must use an unbiased algorithm since the projection that we propose is compatible with the natural gradient, but not necessarily biased estimates thereof.
NAC-Sarsa and INAC are equivalent \emph{biased} discounted-reward natural actor-critic algorithms with per-time-step time complexity linear in the number of features. The former was derived by replacing the LSTD-Q$(\lambda)$ component of NAC-LSTD with Sarsa$(\lambda)$, while the latter is the discounted-reward version of NGAC. Both are similar to NTD, which is a biased average-reward algorithm. The \emph{unbiased} \emph{discounted}-reward form of NAC-Sarsa was recently derived \cite{Thomas2012b}. References to NAC-Sarsa hereafter refer to this unbiased variant. In our case studies we use the \emph{projected natural actor-critic using Sarsa$(\lambda)$} (PNAC-Sarsa), the projected version of the unbiased NAC-Sarsa algorithm.
Notice that the projection, $\Pi_\Theta^{G(\theta_k)}$, as defined in \eqref{eq:blah}, is \emph{not} merely the Euclidean projection back onto $\Theta$. For example, if $\Theta$ is the set of $\theta$ that satisfy $A\theta \leq b$, for some fixed matrix $A$ and vector $b$, then the projection, $\Pi_\Theta^{G(\theta_k)}$, of $y$ onto $\Theta$ is a quadratic program,
\begin{align}
\mbox{minimize } f(\theta)=&-y^\intercal G(\theta_k)\theta +\frac{1}{2}\theta^\intercal G(\theta_k)\theta, \hspace{1cm} \mbox{s.t. }A\theta \leq b.
\end{align}
\vspace{-.1cm}In order to perform this projection, we require an estimate of the average Fisher information matrix, $G(\theta_k)$. If the natural actor-critic algorithm does not already include this (like NAC-LSTD and NAC-Sarsa do not), then an estimate can be generated by selecting $G_0 = \beta I$, where $\beta$ is a positive scalar and $I$ is the identity matrix, and then updating the estimate with
\begin{align}
G_{t+1} = (1-\mu_t) G_t + \mu_t \left ( \frac{\partial}{\partial \theta_k} \log \pi(a_t|s_t,\theta_k) \right ) \left ( \frac{\partial}{\partial \theta_k} \log \pi(a_t|s_t,\theta_k) \right )^\intercal,
\end{align}
where $\{\mu_t\}$ is a stepsize schedule \cite{Bhatnagar2009}. Notice that we use $t$ and $k$ subscripts since many time steps of the MDP may pass between updates to the policy parameters.
\section{Case Study: Functional Electrical Stimulation}
In this case study, we searched for proportional-derivative (PD) gains to control a simulated human arm undergoing FES. We used the Dynamic Arm Simulator 1 (DAS1) \cite{Blana2009}, a detailed biomechanical simulation of a human arm undergoing functional electrical stimulation. In a previous study, a controller created using DAS1 performed well on an actual human subject undergoing FES, although it required some additional tuning in order to cope with biceps spasticity \cite{KathyPD}. This suggests that it is a reasonably accurate model of an ideal arm.
The DAS1 model, depicted in Figure 2a, has state $s_t=(\phi_1, \phi_2, \dot \phi_1, \dot \phi_2, \phi_1^{target}, \phi_2^{target})$, where $\phi_1^{target}$ and $\phi_2^{target}$ are the desired joint angles, and the desired joint angle velocities are zero. The goal is to, during a two-second episode, move the arm from its random initial state to a randomly chosen stationary target. The arm is controlled by providing a stimulation in the interval $[0,1]$ to each of six muscles. The reward function used was similar to that of \citet{KathyPD}, which punishes joint angle error and high muscle stimulation. We searched for locally optimal PD gains using PNAC-Sarsa where the policy was a PD controller with Gaussian noise added for exploration.
Although DAS1 does not model shoulder dislocation, we added safety constraints by limiting the $l_1$-norm of certain pairs of gains. The constraints were selected to limit the forces applied to the humerus. These constraints can be expressed in the form $A \theta \leq b$, where $A$ is a matrix, $b$ is a vector, and $\theta$ are the PD gains (policy parameters). We compared the performance of three algorithms:
\begin{enumerate}
\vspace{-.2cm}\item {\bf NAC}: NAC-Sarsa with no constraints on $\theta$.
\vspace{-.1cm}\item {\bf PNAC}: PNAC-Sarsa using the compatible projection, $\Pi_\Theta^{G(\theta_k)}$.
\vspace{-.1cm}\item {\bf PNAC-E}: PNAC-Sarsa using the Euclidean projection.
\end{enumerate}
\vspace{-.2cm}
Since we are not promoting the use of one natural actor-critic over another, we did not focus on finely tuning the natural actor-critic nor comparing the learning speeds of different natural actor-critics. Rather, we show the importance of the proper projection by allowing PNAC-Sarsa to run for a million episodes (far longer than required for convergence), after which we plot the mean sum of rewards during the last quarter million episodes. Each algorithm was run ten times, and the results averaged and plotted in Figure 2b. Notice that PNAC performs worse than the unconstrained NAC. This happens because NAC leaves the safe region of policy space during its search, and converges to a dangerous policy---one that reaches the goal quickly and with low total muscle force, but which can cause large, short, spikes in muscle forces surrounding the shoulder, which violates our safety constraints. We suspect that PNAC converges to a near-optimal policy within the region of policy space that we have designated as safe. PNAC-E converges to a policy that is worse than that found by PNAC because it uses an incompatible projection.
\begin{figure}[]
\centering
\begin{subfigure}[t]{0.475\textwidth}
\centering
\includegraphics[width=.7\textwidth]{DAS1.png}
\caption*{(Figure 2a) DAS1, the two-joint, six-muscle biomechanical model used. Antagonistic muscle pairs are as follows, listed as (flexor, extensor): monoarticular shoulder muscles (a: anterior deltoid, b: posterior deltoid); monoarticular elbow muscles (c: brachialis, d: triceps brachii (short head)); biarticular muscles (e: biceps brachii, f: triceps brachii (long head)).}
\label{fig:DAS1}
\end{subfigure}
\hspace{.03\textwidth}
\begin{subfigure}[t]{0.475\textwidth}
\centering
\includegraphics[width=1\textwidth]{das1Results.pdf}
\caption*{(Figure 2b) Mean return during the last 250,000 episodes of training using thee algorithms. Standard deviation error bars from the 10 trials are provided. The NAC bar is red to emphasize that the final policy found by NAC resides in the dangerous region of policy space.}
\label{fig:DAS1Data}
\end{subfigure}
\vspace{-.4cm}
\end{figure}
\section{Case Study: uBot Balancing}
In the previous case study, the optimal policy lay outside the designated safe region of policy space (this is common when a single failure is so costly that adding a penalty to the reward function for failure is impractical, since a single failure is unacceptable). We present a second case study in which the optimal policy lies within the designated safe region of policy space, but where an unconstrained search algorithm may enter the unsafe region during its search of policy space (at which point large negative rewards return it to the safe region).
The uBot-5, shown in Figure \ref{fig:uBot}, is an 11-DoF mobile manipulator developed at the University of Massachusetts Amherst \cite{Deegan2010,Kuindersma2009}. During experiments, it often uses its arms to interact with the world. Here, we consider the problem faced by the controller tasked with keeping the robot balanced during such experiments. To allow for results that are easy to visualize in 2D, we use a PD controller that observes only the current body angle, its time derivative, and the target angle (always vertical). This results in the PD controller having only two gains (tunable policy parameters). We use a crude simulation of the uBot-5 with random upper-body movements, and search for the PD gains that minimize a weighted combination of the energy used and the mean angle error (distance from vertical).
We constructed a set of conservative estimates of the region of stable gains, with which the uBot-5 should never fall, and used PNAC-Sarsa and NAC-Sarsa to search for the optimal gains. Each training episode lasted 20 seconds, but was terminated early (with a large penalty) if the uBot-5 fell over. Figure \ref{fig:uBot} (middle) shows performance over 100 training episodes. Using NAC-Sarsa, the PD weights often left the conservative estimate of the safe region, which resulted in the uBot-5 falling over. Figure \ref{fig:uBot} (right) shows one trial where the uBot-5 fell over four times (circled in red). The resulting large punishments cause NAC-Sarsa to quickly return to the safe region of policy space. Using PNAC-Sarsa, the simulated uBot-5 never fell. Both algorithms converge to gains that reside within the safe region of policy space. We selected this example because it shows how, even if the optimal solution resides within the safe region of policy space (unlike the in the previous case study), unconstrained RL algorithms may traverse unsafe regions of policy space during their search.
\begin{figure}
\raisebox{0.8cm}{\includegraphics[width=.18\columnwidth]{uBot.jpg}}
\hspace{.1cm}
\includegraphics[width=.4\columnwidth]{ubot.pdf}
\hspace{-.1cm}
\raisebox{-.2cm}{\includegraphics[width=.3\columnwidth]{uBot2.pdf}} \hfill
\caption{{\bf Left:} uBot-5 holding a ball. {\bf Middle:} Mean (over 20-trials) returns over time using PNAC-Sarsa and NAC-Sarsa on the simulated uBot-5 balancing task. The shaded region depicts standard deviations. {\bf Right:} Trace of the two PD gains, $\theta_1$ and $\theta_2$, from a typical run of PNAC-Sarsa and NAC-Sarsa. A marker is placed for the gains after each episode, and red markers denote episodes where the simulated uBot-5 fell over.}
\label{fig:uBot}
\end{figure}
\section{Summary}
We presented a class of algorithms, which we call \emph{projected natural actor-critics} (PNACs). PNACs are the simple modification of existing natural actor-critic algorithms to include a projection of newly computed policy parameters back onto an allowed set of policy parameters (e.g., those of policies that are known to be safe). We argued that a principled projection is the one that results from viewing natural gradient descent, which is an \emph{unconstrained} algorithm, as a special case of mirror descent, which is a \emph{constrained} algorithm.
We show that the resulting projection is compatible with the natural gradient and gave a simple empirical example that shows why a compatible projection is important. This example also shows how an incompatible projection can result in natural gradient descent converging to a pessimal solution in situations where a compatible projection results in convergence to an optimal solution. We then applied a PNAC algorithm to a realistic constrained control problem with six-dimensional continuous states and actions. Our results support our claim that the use of an incompatible projection can result in convergence to inferior policies. Finally, we applied PNAC to a simulated robot and showed its substantial benefits over unconstrained natural actor-critic algorithms.
\chapter{True Stochastic Gradient Temporal Difference Learning Algorithms}
\label{gtd-chapter}
We now turn to the solution of a longstanding puzzle: how to design a ``true" gradient method for reinforcement learning?
We address long-standing questions in reinforcement learning: (1) Are there any first-order reinforcement
learning algorithms that can be viewed as ``true" stochastic gradient methods?
If there are, what are their objective functions and what are their convergence rates?
(2) What is the general framework for avoiding biased sampling (instead of double-sampling, which is a stringent sampling requirement)
in reinforcement learning? To this end, we introduce a novel primal-dual
splitting framework for reinforcement learning, which shows that the GTD family of algorithms are
true stochastic algorithms with respect to the primal-dual formulation
of the objective functions such as NEU and MSPBE, which facilitates their
convergence rate analysis and regularization. We also propose operator splitting
as a unified framework to avoid bias sampling in reinforcement learning.
We present an illustrative empirical study on simple canonical problems
validating the effectiveness of the proposed algorithms compared with previous
approaches.
\section{Introduction}
First-order temporal difference (TD) learning is a widely used class of techniques
in reinforcement learning. Although least-squares based
temporal difference approaches, such as LSTD \cite{bradtke-barto:LSTD},
LSPE \cite{bertsekas-nedic} and LSPI \cite{lagoudakis:jmlr} perform
well with moderate size problems, first-order temporal difference
learning algorithms scale more gracefully to high dimensional problems.
The initial class of TD methods was known to converge only when samples are drawn
``on-policy". This motivated the development of the gradient TD (GTD) family of methods \cite{FastGradient:2009}. A novel saddle-point framework for sparse regularized GTD was proposed recently \cite{ROTD:NIPS2012}. However, there have been several questions
regarding the current off-policy TD algorithms.
(1) The first is the convergence rate of these algorithms. Although
these algorithms are motivated from the gradient of an objective function
such as MSPBE and NEU, they are not true stochastic gradient methods
with respect to these objective functions, as pointed out in \cite{szepesvari2010algorithms}, which make the convergence rate and error bound analysis difficult, although asymptotic analysis
has been carried out using the ODE approach. (2) The second concern is
regarding acceleration. It is believed that TDC performs the best so far of
the GTD family of algorithms. One may intuitively ask if there are any gradient TD
algorithms that can outperform TDC. (3) The third concern is regarding compactness
of the feasible set $\theta$. The GTD family of algorithms all assume
that the feasible set $\theta$ is unbounded, and if the feasible
set $\theta$ is compact, there is no theoretical analysis and convergence
guarantee. (4) The fourth question is on regularization: although
the saddle point framework proposed in \cite{ROTD:NIPS2012} provides an online regularization framework
for the GTD family of algorithms, termed as RO-TD, it is based on the
inverse problem formulation and is thus not quite explicit. One further
question is whether there is a more straightforward algorithm, e.g,
the regularization is directly based on the MSPBE and NEU objective
functions.
Biased sampling is a well-known problem in reinforcement learning.
Biased sampling is caused by the stochasticity of the policy wherein
there are multiple possible successor states from the current state
where the agent is. If it is a deterministic policy, then there will
be no biased sampling problem. Biased sampling is often caused
by the product of the TD errors, or the product of TD error and the
gradient of TD error w.r.t the model parameter $\theta$. There are
two ways to avoid the biased sampling problem, which can be categorized
into double sampling methods and two-time-scale stochastic approximation
methods.
In this paper, we propose a novel approach to TD algorithm design in reinforcement learning, based on
introducing the {\em proximal splitting} framework \cite{PROXSPLITTING2011}. We show that the GTD family of algorithms are true stochastic gradient descent (SGD) methods, thus making their convergence
rate analysis available. New accelerated off-policy algorithms are
proposed and their comparative study with RO-TD is carried out to show
the effectiveness of the proposed algorithms. We also show that primal-dual
splitting is a unified first-order optimization framework to solve the
biased sampling problem.
Here is a roadmap to the rest of the chapter. Section 2 reviews reinforcement
learning and the basics of proximal splitting formulations and algorithms.
Section 3 introduces a novel problem formulation which we investigate
in this paper. Section 4 proposes a series of new algorithms, demonstrates
the connection with the GTD algorithm family, and also presents accelerated
algorithms. Section 5 presents theoretical analysis of the algorithms.
Finally, empirical results are presented in Section 6 which validate
the effectiveness of the proposed algorithmic framework. Abbreviated technical proofs
of the main theoretical results are provided in a supplementary appendix.
\section{Background}
\subsection{Markov Decision Process and Reinforcement Learning}
In linear value function approximation,
a value function is assumed to lie in the linear span of a basis function
matrix $\Phi$ of dimension $\left|S\right|\times d$, where $d$
is the number of linear independent features. Hence, $V\approx V_{\theta}=\Phi\theta$.
For the $t$-th sample, $\phi_{t}$ (the $t$-th row of $\Phi$),
$\phi'_{t}$ (the $t$-th row of $\Phi'$) are the feature vectors
corresponding to $s_{t},s'_{t}$, respectively. $\theta_{t}$ is the
weight vector for $t$-th sample in first-order TD learning methods,
and ${\delta_{t}}=({r_{t}}+\gamma\phi_{t}^{'T}{\theta_{t}})-\phi_{t}^{T}{\theta_{t}}$
is the temporal difference error. TD learning uses the following update
rule ${\theta_{t+1}}={\theta_{t}}+{\alpha_{t}}{\delta_{t}}{\phi_{t}}$,
where $\alpha_{t}$ is the stepsize. However, TD is only guaranteed
to converge in the on-policy setting, although in many off-policy
situations, it still has satisfactory performance \cite{Kolter:offpolicyTD}.
To this end, Sutton et al. proposed a family of off-policy convergent
algorithms including GTD, GTD2 and TD with gradient correction (TDC).
GTD is a two-time-scale stochastic approximation approach which aims
to minimize the norm of the expected TD update (NEU), which is defined
as
\begin{equation}
{\rm {NEU}}(\theta)=\mathbb{E}{[\delta_{t}(\theta)\phi_{t}]^{T}}\mathbb{E}[\delta_{t}(\theta)\phi_{t}].\label{eq:neu}
\end{equation}
TDC \cite{FastGradient:2009} aims to minimize the mean-square projected
Bellman error (MSPBE) with a similar two-time-scale technique, which
is defined as ${\rm {MSPBE}}(\theta) =$
\begin{equation}
\left\Vert {\Phi\theta-\Pi T(\Phi\theta)}\right\Vert _{\Xi}^{2}={({\Phi^{T}}\Xi(T\Phi\theta-\Phi\theta))^{T}}{({\Phi^{T}}\Xi\Phi)^{-1}}{\Phi^{T}}\Xi(T\Phi\theta-\Phi\theta), \label{eq:mspbe-1}
\end{equation}
where $\Xi$ is a diagonal matrix whose entries $\xi(s)$ are given
by a positive probability distribution over states.
\section{Problem Formulation}
Biased sampling is a well-known problem in reinforcement learning.
Biased sampling is caused by $\mathbb{E}[{\phi_{t}^{'T}}{\phi_{t}^{'}}]$
or $\mathbb{E}[{\phi_{t}^{'}}{\phi_{t}^{'T}}]$, where $\phi_{t}^{'}$
is the feature vector for state $s_{t}^{'}$ in sample $(s_{t},a_{t},r_{t},{s_{t}^{'}})$.
Due to the stochastic nature of the policy, there may be many $s_{t}'$
w.r.t the same $s_{t}$, thus $\mathbb{E}[{\phi_{t}^{'T}}{\phi_{t}^{'}}]$
or $\mathbb{E}[{\phi_{t}^{'}}{\phi_{t}^{'T}}]$ cannot be consistently
estimated via a single sample. This problem hinders the objective
functions to be solved via stochastic gradient descent (SGD) algorithms.
As pointed out in \cite{szepesvari2010algorithms}, although many
algorithms are motivated by well-defined convex objective functions
such as MSPBE and NEU, due to the biased sampling problem, the unbiased
stochastic gradient is impossible to obtain, and thus the algorithms
are not true SGD methods w.r.t. these objective functions.
The biased sampling is often caused by the product of the TD errors,
or the product of TD error and the derivative of TD error w.r.t. the
parameter $\theta$. There are two ways to avoid the biased sampling
problem, which can be categorized into double sampling methods and
stochastic approximation methods. Double sampling, which samples both
$s'$ and $s''$ and thus requires computing $\phi'$ and $\phi''$,
is possible in batch reinforcement learning, but is usually impractical
in online reinforcement learning. The other approach is
stochastic approximation, which introduces a new variable to
estimate the part containing $\phi_{t}^{'}$, thus avoiding the product
of $\phi_{t}^{'}$ and $\phi_{t}^{''}$. Consider, for example, the
NEU objective function in Section (\ref{eq:neu}). Taking the gradient
w.r.t. $\theta$, we have
\begin{equation}
\label{eq:neu-grad}
-\frac{1}{2}{\rm {NEU}}(\theta)=\mathbb{E}[(\phi_{t}-\gamma\phi_{t}'){\phi_{t}^{T}}]\mathbb{E}[\delta_{t}(\theta)\phi_{t}]
\end{equation}
If the gradient can be written as a single expectation value, then
it is straightforward to use a stochastic gradient method, however,
here we have a product of two expectations, and due to the correlation
between $(\phi_{t}-\gamma\phi_{t}'){\phi_{t}^{T}}$ and $\delta_{t}(\theta)\phi_{t}$,
the sampled product is not an unbiased estimate of the gradient. In
other words, $\mathbb{E}[(\phi_{t}-\gamma\phi_{t}'){\phi_{t}^{T}}]$
and $\mathbb{E}[\delta_{t}(\theta)\phi_{t}]$ can be directly sampled,
yet $\mathbb{E}[(\phi_{t}-\gamma\phi_{t}'){\phi_{t}^{T}}]\mathbb{E}[\delta_{t}(\theta)\phi_{t}]$
can not be directly sampled. To tackle this, the GTD algorithm uses
the two-time-scale stochastic approximation method by introducing
an auxiliary variable $w_{t}$, and thus the method is not a true
stochastic gradient method w.r.t. ${\rm {NEU}}(\theta)$ any more.
This auxiliary variable technique is also used in \cite{ZHIWEI2014}.
The other problem for first-order reinforcement learning algorithms
is that it is difficult to define the objective functions, which is
also caused by the biased sampling problem. As pointed out in \cite{szepesvari2010algorithms},
although the GTD family of algorithms are derived from the gradient w.r.t.
the objective functions such as MSPBE and NEU, because of the biased-sampling
problem, these algorithms cannot be formulated directly as SGD methods
w.r.t. these objective functions.
In sum, due to biased sampling, the RL objective functions cannot
be solved via a stochastic gradient method, and it is also difficult
to find objective functions of existing first-order reinforcement learning
algorithms. Thus, there remains a large gap between first-order reinforcement
learning algorithms and stochastic optimization, which we now show how to bridge.
\section{Algorithm Design}
In what follows, we build on the operator splitting methods introduced in Section~\ref{pd-splitting}, which should be reviewed before reading
the section below.
\subsection{NEU Objective Function}
The primal-dual formulation of the NEU defined in Section (\ref{eq:neu})
is as follows:
\begin{equation}
\mathop{\min}\limits _{\theta\in X}\left({\frac{1}{2}{\rm {NEU}}(\theta)+h(\theta)}\right)=\mathop{\min}\limits _{\theta\in X}\mathop{\max}\limits _{y}\left({\langle {\Phi^{T}}\Xi(R+\gamma{\Phi^{'}}\theta-\Phi\theta),y\rangle -\frac{1}{2}||y||_{2}^{2}+h(\theta)}\right)\label{eq:lower-neu}
\end{equation}
We have $K(\theta)={\Phi^{T}}\Xi(R+\gamma{\Phi^{'}}\theta-\Phi\theta)$
, and $F(\cdot)=\frac{1}{2}||\cdot||_{2}^{2}$ , thus the Legendre
transform is ${F^{*}}(\cdot)=F(\cdot)=\frac{1}{2}||\cdot||_{2}^{2}$.
Thus the update rule is
\begin{equation}
\begin{array}{l}
{y_{t+1}}={y_{t}}+{\alpha_{t}}({\delta_{t}}{\phi_{t}}-{y_{t}}){\rm {,\,}}
{\theta _{t + 1}} = {\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}\left( {{\theta _t} + {\alpha _t}({\phi _t} - \gamma \phi _t^\prime )(y_t^T{\phi _t})} \right)
\end{array}\label{eq:gtd}
\end{equation}
Note that if $h(\theta)=0$ and $X={\mathbb{R}^{d}}$, then we will
have the GTD algorithm proposed in \cite{Sutton:GTD1:2008}.
\subsection{MSPBE Objective Function }
Based on the definition of MSPBE in Section (\ref{eq:neu}), we
can reformulate MSPBE as
\begin{equation}
{\rm {MSPBE}}(\theta)=||{\Phi^{T}}\Xi(T{V_{\theta}}-{V_{\theta}})||_{{{({\Phi^{T}}\Xi\Phi)}^{-1}}}^{2}\label{eq:mspbe-1-1}
\end{equation}
The gradient of MSPBE is correspondingly computed as
\begin{equation}
-\frac{1}{2}{\rm {MSPBE}}(\theta)=\mathbb{E}[({\phi_{t}}-\gamma{\phi_{t}^{'}})\phi_{t}^{T}]\mathbb{E}{[{\phi_{t}}\phi_{_{t}}^{T}]^{-1}}\mathbb{E}[{\delta_{t}}(\theta){\phi_{t}}]\label{eq:mspbe-grad}
\end{equation}
As opposed to computing the NEU gradient, computing Equation
(\ref{eq:mspbe-grad}) involves computing the inverse matrix $\mathbb{E}{[{\phi_{t}}\phi_{_{t}}^{T}]^{-1}}$,
which imposes extra difficulty. To this end, we propose another primal-dual
splitting formulation with weighted Euclidean norm as follows,
\begin{equation}
\mathop{\min}\limits _{x\in X}\frac{1}{2}||x||_{{M^{-1}}}^{2}=\mathop{\min}\limits _{x\in X}\mathop{\max}\limits _{w}\langle x,w \rangle -\frac{1}{2}||w||_{M}^{2}\label{eq:primdual-mspbe}
\end{equation}
where $M={\Phi^{T}}\Xi\Phi$,
and the dual variable is denoted as $w_{t}$ to differentiate it from
$y_{t}$ used for the NEU objective function. Then we have
\begin{equation}
\mathop{\min}\limits _{\theta\in X}\frac{1}{2}{\rm {MSPBE}}(\theta)+h(\theta)=\mathop{\min}\limits _{\theta\in X}\mathop{\max}\limits _{w}\langle {{\Phi^{T}}\Xi(R+\gamma{\Phi^{'}}\theta-\Phi\theta),w}\rangle -\frac{1}{2}||w||_{M}^{2}+h(\theta)\label{eq:lower-mspbe}
\end{equation}
Note that the nonlinear
convex $F(\cdot)=\frac{1}{2}||\cdot||_{M^{-1}}^{2}$ , and thus the
Legendre transform is ${F^{*}}(\cdot)=\frac{1}{2}||\cdot||_{M}^{2}$.
We can see that by using the primal-dual splitting formulation, computing
the inverse matrix $M^{-1}$ is avoided. Thus the update rule
is as follows:
\begin{equation}
\begin{array}{l}
{w_{t+1}}={w_{t}}+{\alpha_{t}}({\delta_{t}}-\phi_{_{t}}^{T}{w_{t}}){\phi_{t}}{\rm {,\,\mathbf{}}}\mbox{{\ensuremath{\theta_{t+1}}} = pro{\ensuremath{x_{{\alpha_{t}}h}}}\ensuremath{\left({{\theta_{t}}+{\alpha_{t}}({\phi_{t}}-\gamma\phi_{t}^{\prime})(w_{t}^{T}{\phi_{t}})}\right)}}\end{array}\label{eq:gtd2}
\end{equation}
Note that if $h(\theta)=0$ and $X={\mathbb{R}^{d}}$, then we will
have the GTD2 algorithm proposed in \cite{FastGradient:2009}. It
is also worth noting that the TDC algorithm seems not to have an explicit proximal
splitting representation, since it incorporates $w_{t}(\theta)=\mathbb{E}{[{\phi_{t}}\phi_{_{t}}^{T}]^{-1}}\mathbb{E}[{\delta_{t}}(\theta){\phi_{t}}]$
into the update of $\theta_{t}$, a quasi-stationary condition which
is commonly used in two-time-scale stochastic approximation approaches.
An intuitive answer to the advantage of TDC over GTD2 is that
the TDC update of $\theta_{t}$ can be considered as incorporating
the prior knowledge into the update rule: for a stationary $\theta_{t}$,
if the optimal $w_{t}(\theta_{t})$ (termed as $w_{t}^{*}(\theta_{t})$)
has a closed-form solution or is easy to compute, then incorporating
this $w_{t}^{*}(\theta_{t})$ into the update rule tends to accelerate
the algorithm's convergence performance. For the GTD2 update in Equation
(\ref{eq:gtd2}), note that there is a sum of two terms where $w_{t}$
appears: which are $({\phi_{t}}-\gamma\phi_{t}^{\prime})(w_{t}^{T}{\phi_{t}})={\phi_{t}}(w_{t}^{T}{\phi_{t}})-\gamma\phi_{t}^{\prime}(w_{t}^{T}{\phi_{t}})$. Replacing $w_{t}$ in the first term with $w_{t}^{*}(\theta)=\mathbb{E}{[{\phi_{t}}\phi_{_{t}}^{T}]^{-1}}\mathbb{E}[{\delta_{t}}(\theta){\phi_{t}}]$,
we have the update rule as follows
\begin{equation}
\begin{array}{l}
{w_{t+1}}={w_{t}}+{\alpha_{t}}({\delta_{t}}-\phi_{_{t}}^{T}{w_{t}}){\phi_{t}}
{\;,\;}
\mbox{{\ensuremath{\theta_{t+1}}} = pro{\ensuremath{x_{{\alpha_{t}}h}}}\ensuremath{\left({{\theta_{t}}+{\alpha_{t}}({\phi_{t}}-\gamma\phi_{t}^{\prime})(\phi_{t}^{T}{w_{t}})}\right)}}
\end{array}
\end{equation}
Note that if $h(\theta)=0$ and $X={\mathbb{R}^{d}}$, then we will
have TDC algorithm proposed in \cite{FastGradient:2009}. Note that
this technique does not have the same convergence guarantee as the original
objective function. For example, if we use a similar trick on the GTD update
with the optimal $y_{t}(\theta_{t})$ (termed as $y_{t}^{*}(\theta_{t})$)
where $y_{t}^{*}(\theta)=\mathbb{E}[{\delta_{t}}(\theta){\phi_{t}}]$,
then we can have
\begin{equation}
\theta_{t+1}={\rm pro}{\ensuremath{{\rm x}_{{\alpha_{t}}h}}}\ensuremath{\left({\theta_{t}}+{\alpha_{t}}{\delta_{t}}({\phi_{t}}-\gamma\phi_{t}^{'})\right)}
\end{equation}
which is the update rule of residual gradient \cite{Baird:ResidualAlgorithms1995},
and is proven not to converge to NEU any more.%
\footnote{It converges to mean-square TD error (MSTDE), as proven in \cite{maei2011gradient}.%
}
\section{Accelerated Gradient Temporal Difference Learning Algorithms}
In this section we will discuss the acceleration of GTD2 and TDC.
The acceleration of GTD is not discussed due to space consideration,
which is similar to GTD2. A comprehensive
overview of the convergence rate of different approaches to stochastic
saddle-point problems is given in \cite{chen2013optimal}. In this section we present accelerated
algorithms based on the Stochastic Mirror-Prox (SMP) Algorithm \cite{sra2011optimization,juditsky2008solving}.
Algorithm \ref{alg:GTD2mp}, termed as GTD2-MP, is accelerated GTD2
with extragradient. Algorithm \ref{alg:TDCmp}, termed as TDC-MP,
is accelerated TDC with extragradient.
\begin{algorithm}
\caption{Algorithm Template}
\label{alg:tdneu}
Let $\pi$ be some fixed policy of an MDP $M$, $\Phi$ be some fixed basis.
\begin{algorithmic}[1]
\REPEAT
\STATE Compute $\ensuremath{{{\phi_{t}},{\phi_{t}}'}}$ and TD error ${\delta_{t}}={r_{t}}+\gamma{\phi_{t}}^{\prime T}{\theta_{t}}-\phi_{t}^{T}{\theta_{t}}$
\STATE Compute $\ensuremath{\theta_{t+1},w_{t+1}}$ according to each algorithm update rule\
\UNTIL {$t=N$};
\STATE Compute primal average ${\bar{\theta}_{N}}=\frac{1}{N}\sum\limits _{i=1}^{N}{\theta_{i}},{\bar{w}_{N}}=\frac{1}{N}\sum\limits _{i=1}^{N}{w_{i}}$
\end{algorithmic}
\end{algorithm}
\begin{algorithm}
\caption{GTD2-MP}
\label{alg:GTD2mp}
\begin{enumerate}
\item ${w_{t+\frac{1}{2}}}={{w_{t}}+{\beta_{t}}({\delta_{t}}-\phi_{t}^{T}{w_{t}}){\phi_{t}}},\\ \;{\theta_{t+\frac{1}{2}}}={{\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}}\left({{\theta_{t}}+{\alpha_{t}}({\phi_{t}}-\gamma{\phi_{t}}^{\prime})(\phi_{t}^{T}{w_{t}})}\right)$
\item ${\delta_{t+\frac{1}{2}}}={r_{t}}+\gamma{\phi_{t}}^{\prime T}{\theta_{t+\frac{1}{2}}}-\phi_{t}^{T}{\theta_{t+\frac{1}{2}}}$
\item $\begin{array}{l}
{w_{t+1}}={w_{t}}+{\beta_{t}}({\delta_{t+\frac{1}{2}}}-\phi_{t}^{T}{w_{t+\frac{1}{2}}}){\phi_{t}}
{\;,\;}\\
{\theta_{t+1}}={{\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}}\left({{\theta_{t}}+{\alpha_{t}}({\phi_{t}}-\gamma{\phi_{t}}^{\prime})(\phi_{t}^{T}{w_{t+\frac{1}{2}}})}\right)
\end{array}
$\end{enumerate}
\end{algorithm}
\begin{algorithm}
\caption{TDC-MP}
\label{alg:TDCmp}
\begin{enumerate}
\item ${w_{t+\frac{1}{2}}}={{w_{t}}+{\beta_{t}}({\delta_{t}}-\phi_{t}^{T}{w_{t}}){\phi_{t}}},\\ \;{\theta_{t+\frac{1}{2}}}={{\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}}\left({{\theta_{t}}+{\alpha_{t}}\delta_{t}{\phi_{t}}-{\alpha_{t}}\gamma{\phi_{t}}^{\prime}(\phi_{t}^{T}{w_{t}})}\right)$
\item ${\delta_{t+\frac{1}{2}}}={r_{t}}+\gamma{\phi_{t}}^{\prime T}{\theta_{t+\frac{1}{2}}}-\phi_{t}^{T}{\theta_{t+\frac{1}{2}}}$
\item $\begin{array}{l}
{w_{t+1}}={w_{t}}+{\beta_{t}}({\delta_{t+\frac{1}{2}}}-\phi_{t}^{T}{w_{t+\frac{1}{2}}}){\phi_{t}}\;
{\;,\;}\\
{\theta_{t+1}}={{\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}}\left({{\theta_{t}}+{\alpha_{t}}{\delta_{t+\frac{1}{2}}}{\phi_{t}}-{\alpha_{t}}\gamma{\phi_{t}}^{\prime}(\phi_{t}^{T}{w_{t+\frac{1}{2}}})}\right)
\end{array}$\end{enumerate}
\end{algorithm}
\section{Theoretical Analysis}
In this section, we discuss the convergence rate and error bound of
GTD, GTD2 and GTD2-MP.
\subsection{Convergence Rate}
\textbf{Proposition 1} The convergence rates of the GTD/GTD2
algorithms with primal average are $O(\frac{{{L_{{F^{*}}}}+{L_{K}}+\sigma}}{{\sqrt{N}}})$,
where ${L_{K}}=||{\Phi^{T}}\Xi(\Phi-\gamma{\Phi^{'T}})|{|^{2}}$,
for GTD, ${L_{{F^{*}}}}=1$ and for GTD2, ${L_{{F^*}}} = ||{\Phi ^T}\Xi \Phi |{|_2}$, $\sigma$ is defined in the Appendix due to space limitations.
Now we consider the convergence rate of GTD2-MP.
\textbf{Proposition 2 }The convergence rate of the GTD2-MP algorithm
is $O(\frac{{{L_{{F^{*}}}}+{L_{K}}}}{N}+\frac{\sigma}{{\sqrt{N}}})$.
See supplementary materials for an abbreviated proof.
\textbf{Remark}: The above propositions imply that when the noise level is low, the GTD2-MP algorithm is able to converge at the rate of $O(\frac{1}{N})$, whereas the convergence rate of GTD2 is $O(\frac{1}{{\sqrt N }})$. However, when the noise level is high, both algorithms' convergence rates reduce to $O(\frac{\sigma }{{\sqrt N }})$.
\subsection{Value Approximation Error Bound}
\label{eq:mspbe}
\textbf{Proposition 3}: For GTD/GTD2, the prediction error of $||V-{V_{\theta}}||$
is bounded by $||V-{V_{\theta}}|{|_{\infty}}\le\frac{{L_{\phi}^{\Xi}}}{{1-\gamma}}\cdot O\left({\frac{{{L_{{F^{*}}}}+{L_{K}}+\sigma}}{{\sqrt{N}}}}\right)$
; For GTD2-MP, it is bounded by $||V-{V_{\theta}}|{|_{\infty}}\le\frac{{L_{\phi}^{\Xi}}}{{1-\gamma}}\cdot O\left({\frac{{{L_{{F^{*}}}}+{L_{K}}}}{N}+\frac{\sigma}{{\sqrt{N}}}}\right)$,
where $L_{\phi}^{\Xi}={\max_{s}}||{({\Phi^{T}}\Xi\Phi)^{-1}}\phi(s)|{|_{1}}$.
\textbf{Proof}: see Appendix.
\subsection{Related Work}
Here we will discuss previous related work. To the best of our knowledge,
the closest related work is the RO-TD algorithm, which first introduced
the convex-concave saddle-point framework to regularize the TDC family
of algorithms. \footnote{Although only regularized TDC was proposed in \cite{ROTD:NIPS2012},
the algorithm can be easily extended to regularized GTD and GTD2.}
The major difference is that RO-TD is motivated by the linear inverse problem formulation of TDC algorithm and uses its dual norm representation as the objective function, which does not explore the auxiliary variable $w_t$. In contrast, by introducing the operator splitting framework, we demonstrate that the GTD family of algorithms can be nicely explained as a ``true'' SGD approach, where the auxiliary variable $w_t$ has a nice explanation.
Another interesting question is whether ADMM is suitable for the operator
splitting algorithm here. Let's take NEU for example. The ADMM formulation
is as follows, where we assume $K(\theta ) = K\theta$ for simplicity, and other scenarios can be derived similarly,
\begin{equation}
\mathop{\min}\limits _{\theta,z}\left({F(z)+h(\theta)}\right){\rm {s}}.{\rm {t}}.{\rm {}}z=K\theta\label{eq:admm}
\end{equation}
The update rule is as follows, where $\alpha_{t}$ is the stepsize
\begin{equation}
\begin{array}{l}
{\theta_{t+1}}=\arg\mathop{\min}\limits _{\theta}\left({h(\theta)+\left\langle {{y_{t}},K\theta-{z_{t}}}\right\rangle +\frac{1}{2}||K\theta-{z_{t}}|{|^{2}}}\right)\\
{z_{t+1}}=\arg\mathop{\min}\limits _{z}\left({F(z)+\left\langle {{y_{t}},K{\theta_{t+1}}-z}\right\rangle +\frac{1}{2}||K{\theta_{t+1}}-z|{|^{2}}}\right)\\
{y_{t+1}}={y_{t}}+\alpha_{t}(K{\theta_{t+1}}-{z_{t+1}})
\end{array}\label{eq:admm-update}
\end{equation}
At first glance the operator of $F(\cdot)$ and $K\theta$ seem to
be split, however, if we compute the closed-form update rule of $\theta_{t}$,
we can see that the update of $\theta_{t}$ includes ${({K^{T}}K)^{-1}}$,
which involves both biased-sampling and computing the inverse matrix,
thus regular ADMM does not seem to be practical for this first-order
reinforcement learning setting. However, using the pre-conditioning technique
introduced in \cite{esser2010general}, ADMM can be reduced to the
primal-dual splitting method as pointed out in \cite{POCK2011SADDLE}.
\section{Experimental Study}
\subsection{Off-Policy Convergence: Baird Example}
The Baird example is a well-known example where TD diverges and TDC converges. The stepsizes are set to be constants where ${\beta _t} = \mu {\alpha _t}$ as shown in Figure \ref{fig:star}. From Figure \ref{fig:star}, we can see that GTD2-MP
and TDC-MP have a significant advantage over the GTD2 and TDC algorithms
wherein both the MSPBE and the variance are substantially reduced.
\begin{figure}
\centering
\begin{minipage}{1.15\textwidth}
\includegraphics[width= .45\textwidth, height=2.25in]{baird_1.pdf}
\includegraphics[width= .45\textwidth, height=2.25in]{baird_tdc.pdf}
\end{minipage}
\caption{Off-Policy Convergence Comparison}
\label{fig:star}
\end{figure}
\subsection{Regularization Solution Path: Two-State Example}
Now we consider the two-state MDP in \cite{DantzigRL:2012}. The transition
matrix and reward vector are $[0,1;0,1]$ and $R = {[0, - 1]^T},\gamma=0.9$, and a one-feature basis $\Phi={[1,2]^{T}}$. The objective function are
$\theta = \arg \mathop {\min }\limits_\theta \left( {\frac{1}{2}L(\theta ) + \rho ||\theta |{|_1}} \right)$,
where $L(\theta)$ is ${{\rm {NEU}}(\theta)}$ and ${{\rm {MSPBE}}(\theta)}$.
The objective functions are termed as $l_{1}$-NEU and $l_{1}$-MSPBE
for short. In Figure \ref{fig:2state}, both $l_{1}$-NEU and $l_{1}$-MSPBE have well-defined solution paths w.r.t $\rho$, whereas Lasso-TD may
have multiple solutions if the $P$-matrix condition is not satisfied
\cite{Kolter09LARSTD}.
\begin{figure}
\centering
\begin{minipage}{1\textwidth}
\includegraphics[width= .32\textwidth, height=1.5in]{lassotd.pdf}
\includegraphics[width= .32\textwidth, height=1.5in]{mspbe.pdf}
\includegraphics[width= .32\textwidth, height=1.5in]{neu.pdf}
\end{minipage}
\caption{Solution Path Comparison}
\label{fig:2state}
\end{figure}
\subsection{On-Policy Performance: $400$-State Random MDP}
In this experiment we compare the on-policy performance of the four algorithms. We use the random generated MDP with $400$ states and $10$ actions in \cite{dann2014tdsurvey}. Each state is represented by a feature vector with $201$ features, where $200$ features are generated by sampling from a uniform distribution the $201$-th feature is a constant. The stepsizes are set to be constants where ${\beta _t} = \mu {\alpha _t}$ as shown in Figure \ref{fig:400state}. The parameters of each algorithm are chosen via comparative studies similar to \cite{dann2014tdsurvey}.
The result is shown in Figure \ref{fig:400state}. The results for each algorithm are averaged on $100$ runs, and the parameters of each algorithm are chosen via experiments. TDC shows high variance and chattering effect of MSPBE curve on this domain.
Compared with GTD2, GTD2-M1P is able to reduce the MSPBE significantly. Compared with TDC, TDC-MP not only reduces the MSPBE, but also the variance and the ''chattering" effect.
\begin{figure}[h]
\centering{}\includegraphics[width=.8\textwidth,height=2.5in]{disc_on_tdc_mp.pdf}
\caption{Comparison of $400$-State Random MDP}
\label{fig:400state}
\end{figure}
\section{Summary}
\label{gtd-summary}
This chapter shows that the GTD/GTD2 algorithms are true stochastic gradient methods {\em w.r.t. the primal-dual formulation of their corresponding objective functions, which enables their convergence rate analysis and
regularization}. Second, it proposes operator splitting as a broad framework to solve the biased-sampling
problem in reinforcement learning. Based on the unified primal-dual splitting framework, it also proposes
accelerated algorithms with both rigorous theoretical analysis and illustrates their
improved performance w.r.t. previous methods. Future research is ongoing to explore other operator splitting techniques
beyond primal-dual splitting as well as incorporating random projections \cite{lstdrp:nips2010}, and investigating kernelized algorithms \cite{barreto2013kernelrl,taylor2009kernelrl}. Finally, exploring the convergence rate of the TDC algorithm is also important
and interesting.
\chapter{Variational Inequalities: The Emerging Frontier of Machine Learning}
\label{vis}
This paper describes a new framework for reinforcement learning based on {\em primal dual} spaces connected by a Legendre transform. The ensuing theory yields surprising and beautiful solutions to several important questions that have remained unresolved: (i) how to design reliable, convergent, and stable reinforcement learning algorithms (ii) how to guarantee that reinforcement learning satisfies pre-specified ``safety" guarantees, and remains in a stable region of the parameter space (iv) how to design ``off-policy" TD-learning algorithms in a reliable and stable manner, and finally, (iii) how to integrate the study of reinforcement learning into the rich theory of stochastic optimization. In this paper, we gave detailed answers to all these questions using the powerful framework of {\em proximal operators}. The single most important idea that emerges is the use of {\em primal dual spaces} connected through the use of a {\em Legendre} transform. This allows temporal-difference updates to occur in dual spaces, allowing a variety of important technical advantages. The Legendre transform, as we show, elegantly generalizes past algorithms for solving reinforcement learning problems, such as {\em natural gradient} methods, which we show relate closely to the previously unconnected framework of {\em mirror descent} methods. Equally importantly, proximal operator theory enables the systematic development of {\em operator splitting} methods that show how to safely and reliably decompose complex products of gradients that occur in recent variants of gradient-based temporal-difference learning. This key technical contribution makes it possible to finally show to design ``true" stochastic gradient methods for reinforcement learning. Finally, Legendre transforms enable a variety of other benefits, including modeling sparsity and domain geometry. Our work builds extensively on recent work on the convergence of saddle-point algorithms, and on the theory of {\em monotone operators} in Hilbert spaces, both in optimization and for variational inequalities. The latter represents possibly the most exciting future research direction, and we give a more detailed description of this ongoing research thrust.
\section{Variational Inequalities}
Our discussion above has repeatedly revolved around the fringes of variational inequality theory. Methods like extragradient \cite{extragradient:1976} and the mirror-prox algorithm were originally proposed to solve variational inequalities and related saddle point problems. We are currently engaged in redeveloping the proposed ideas more fully within the fabric of variational inequality (VI). Accordingly, we briefly describe the framework of VIs, and give the reader a brief tour of this fascinating extension of the basic underlying framework of optimization. We lack the space to do a thorough review. That is the topic of another monograph to be published at a later date, and several papers on this topic are already under way.
At the dawn of a new millennium, the Internet dominates our economic, intellectual and social lives. The concept of equilibrium plays a key role in understanding not only the Internet, but also other networked systems, such as human migration \cite{nagurney:migration}, evolutionary dynamics and the spread of infectious diseases \cite{novak:book}, and social networks \cite{kleinberg:text}. Equilibria are also a central idea in game theory \cite{fudenberg-levine:book,algorithmic-game-theory}, economics \cite{econ-text}, operations research \cite{puterman}, and many related areas. We are currently exploring two powerful mathematical tools for the study of equilibria -- variational inequalities (VIs) and projected dynamical systems (PDS) \cite{nagurney:vibook,nagurney:pdsbook} -- in developing a new machine learning framework for solving equilibrium problems in a rich and diverse range of practical applications. As Figure~\ref{vi-problems} illustrates, finite-dimensional VIs provide a mathematical framework that unifies many disparate equilibrium problems of significant importance, including (convex) optimization, equilibrium problems in economics, game theory and networks, linear and nonlinear complementarity problems, and solutions of systems of nonlinear equations.
\begin{figure}[h]
\begin{center}
\begin{minipage}[t]{0.85\textwidth}
\includegraphics[width=\textwidth,height=3in]{vi-problems.pdf}
\end{minipage} \hspace{0.2in}
\end{center}
\vskip -0.2in
\caption{A variety of real-world problems can be modeled as solving variational inequalities.}
\label{vi-problems}
\end{figure}
Variational inequalities (VIs), in the infinite-dimensional setting, were originally proposed by Hartman and Stampacchia \cite{hartman-stampacchia:acta} in the mid-1960s in the context of solving partial differential equations in mechanics. Finite-dimensional VIs rose in popularity in the 1980s partly as a result of work by Dafermos \cite{dafermos}. who showed that the traffic network equilibrium problem could be formulated as a finite-dimensional VI. This advance inspired much follow-on research, showing that a variety of equilibrium problems in economics, game theory, sequential decision-making etc. could also be formulated as finite-dimensional VIs -- the books by Nagurney \cite{nagurney:vibook} and Facchinei and Pang \cite{facchinei-pang:vi} provide a detailed introduction to the theory and applications of finite-dimensional VIs. Projected dynamical systems (PDS) \cite{nagurney:pdsbook} are a class of ordinary differential equations (ODEs) with a discontinuous right-hand side. Associated with every finite-dimensional VI is a PDS, whose stationary points are the solutions of the VI. While VIs provide a static analysis of equilibria, PDS enable a microscopic examination of the dynamic processes that lead to or away from stable equilibria. . There has been longstanding interest in AI in the development of gradient-based learning algorithms for finding Nash equilibria in multiplayer games, e.g. \cite{bowling:aij,fudenberg-levine:book,singh:uai2000}. A gradient method for finding Nash equilibria can be formalized by a set of ordinary differential equations, whose phase space portrait solution reveals the dynamical process of convergence to an equilibrium point, or lack thereof. A key complication in this type of analysis is that the classical dynamical systems approach does not allow incorporating constraints on values of variables, which are omnipresent in equilibria problems, not only in games, but also in many other applications in economics, network flow, traffic modeling etc. In contrast, the right-hand side of a PDS is a discontinuous projection operator that allows enabling constraints to be modeled.
One of the original algorithms for solving finite-dimensional VIs is the {\em extragradient method} proposed by Korpelevich \cite{korpelevich}. It has been applied to structured prediction models in machine learning by Taskar et al. \cite{taskar:jmlr}. Bruckner et al. \cite{bruckner:jmlr} use a modified extragradient method for solving the spam filtering problem modeled as a prediction game. We are developing a new family of extragradient-like methods based on well-known numerical methods for solving ordinary differential equations, specifically the {\em Runge Kutta} method \cite{numerical-recipes}. In optimization, the extragradient algorithm was generalized to the non-Euclidean case by combining it with the mirror-descent method \cite{nemirovski-yudin:book}, resulting in the so-called ``mirrror-prox" algorithm \cite{mirror-prox, mirror-prox2}. We have extended the mirror-prox method by combining it with Runge-Kutta methods for solving high-dimensional VI problems over the simplex and other spaces. We show the enhanced performance of Runge-Kutta extragradient methods on a range of benchmark variational inequalities drawn from standard problems in the optimization literature.
\subsection{Definition}
\begin{figure}[t]
\begin{center}
\begin{minipage}[t]{0.75\textwidth}
\includegraphics[width=\textwidth,height=1.5in]{vi-geom.pdf}
\end{minipage}
\end{center}
\caption{This figure provides a geometric interpretation of the variational inequality $VI(F,K)$. The mapping $F$ defines a vector field over the feasible set $K$ such that at the solution point $x^*$, the vector field $F(x^*)$ is directed inwards at the boundary, and $-F(x^*)$ is an element of the normal cone $C(x^*)$ of $K$ at $x^*$.}
\label{vi-geom}
\end{figure}
The formal definition of a VI as follows:\footnote{Variational problems can be defined more abstractly in Hilbert spaces. We confine our discussion to $n$-dimensional Euclidean spaces.}
\begin{definition}
The finite-dimensional variational inequality problem VI(F,K) involves finding a vector $x^* \in K \subset \mathbb{R}^n$ such that
\[ \langle F(x^*), x - x^* \rangle \geq 0, \ \forall x \in K \]
where $F: K \rightarrow \mathbb{R}^n$ is a given continuous function and $K$ is a given closed convex set, and $\langle ., . \rangle$ is the standard inner product in $\mathbb{R}^n$.
\end{definition}
Figure~\ref{vi-geom} provides a geometric interpretation of a variational inequality. \footnote{In Figure~\ref{vi-geom}, the normal cone $C(x^*)$ at the vector $x^*$ of a convex set $K$ is defined as $C(x^*) = \{y \in \mathbb{R}^n | \langle y, x - x^* \rangle \leq 0, \forall x \in K \}$. } The following general result characterizes when solutions to VIs exist:
\begin{theorem}
Suppose $K$ is compact, and that $F: K \rightarrow \mathbb{R}^n$ is continuous. Then, there exists a solution to VI($F,K$).
\end{theorem}
As Figure~\ref{vi-geom} shows, $x^*$ is a solution to $VI(F,K)$ if and only if the angle between the vectors $F(x^*)$ and $x - x^*$, for any vector $x \in K$, is less than or equal to $90^0$. To build up some intuition, the reduction of a few well-known problems to a VI is now provided.
\begin{theorem}
Let $x^*$ be a solution to the optimization problem of minimizing a continuously differentiable function $f(x)$, subject to $x \in K$, where $K$ is a closed and convex set. Then, $x^*$ is a solution to $VI(\nabla f, K)$, such that $\langle \nabla f(x^*), x - x^* \rangle \geq 0, \ \forall x \in K$.
\end{theorem}
{\bf Proof:} Define $\phi(t) = f(x^* + t (x - x^*))$. Since $\phi(t)$ is minimized at $t=0$, it follows that $0 \leq \phi'(0) = \langle \nabla f(x^*), x - x^* \rangle \geq 0, \ \forall x \in K$, that is $x^*$ solves the VI.
\begin{theorem}
If $f(x)$ is a convex function, and $x^*$ is the solution of $VI(\nabla f, K)$, then $x^*$ minimizes $f$.
\end{theorem}
{\bf Proof:} Since $f$ is convex, it follows that any tangent lies below the function, that is $f(x) \geq f(x^*) + \langle \nabla f(x^*), x - x^* \rangle, \ \forall x \in K$. But, since $x^*$ solves the VI, it follows that $f(x^*)$ is a lower bound on the value of $f(x)$ everywhere, or that $x^*$ minimizes $f$.
A rich class of problems called {\bf complementarity problems} (CPs) also can be reduced to solving a VI. When the feasible set $K$ is a cone, meaning that if $x \in K$, then $\alpha x \in K, \alpha \geq 0$, then the VI becomes a CP.
\begin{definition}
Given a cone $K \subset \mathbb{R}^n$, and a mapping $F: K \rightarrow \mathbb{R}^n$, the complementarity problem CP(F,K) is to find an $x \in K$ such that $F(x) \in K^*$, the dual cone to $K$, and $\langle x, F(x) \rangle \geq 0$. \footnote{Given a cone $K$, the dual cone $K^*$ is defined as $K^* = \{ y \in \mathbb{R}^n | \langle y, x \rangle \geq 0, \forall x \in K \}$.}
\end{definition}
A number of special cases of CPs are important. The nonlinear complementarity problem (NCP) is to find $x^* \in \mathbb{R}^n_+$ (the non-negative orthant) such that $F(x^*) \geq 0$ and $\langle F(x^*), x^* \rangle = 0$. The solution to an NCP and the corresponding $VI(F, \mathbb{R}^n_+)$ are the same, showing that NCPs reduce to VIs. In an NCP, whenever the mapping function $F$ is affine, that is $F(x) = Mx + b$, where $M$ is an $n \times n$ matrix, then the corresponding NCP is called a linear complementarity problem (LCP) \cite{murty:lcpbook}. Recent work on learning sparse models using $L_1$ regularization has exploited the fact that the standard LASSO objective \cite{lasso} of $L_1$ penalized regression can be reduced to solving an LCP \cite{lasso-lcp}. This reduction to LCP has been used in recent work on sparse value function approximation as well in a method called LCP-TD \cite{jeff:lcp-rl}.
A final crucial property of VIs is that they can be formulated as finding {\bf fixed points}.
\begin{theorem}
\label{projthm}
The vector $x^*$ is the solution of VI(F,K) if and only if, for any $\gamma > 0$, $x^*$ is also a fixed point of the map $x^* = \Pi_K(x^* - \gamma F(x^*))$,
where $\Pi_K$ is the projector onto convex set $K$.
\end{theorem}
In terms of the geometric picture of a VI illustrated in Figure~\ref{vi-geom}. this property means that the solution of a VI occurs at a vector $x^*$ where the vector field $F(x^*)$ induced by $F$ on $K$ is normal to the boundary of $K$ and directed inwards, so that the projection of $x^* - \gamma F(x^*)$ is the vector $x^*$ itself. This property forms the basis for the projection class of methods that solve for the fixed point.
\subsection{Equilibrium Problems in Game Theory}
The VI framework provides a mathematically elegant approach to model equilibrium problems in game theory \cite{fudenberg-levine:book,algorithmic-game-theory}. A {\em Nash game} consists of $m$ players, where player $i$ chooses a strategy $x_i$ belonging to a closed convex set $X_i \subset \mathbb{R}^n$. After executing the joint action, each player is penalized (or rewarded) by the amount $F_i(x_1, \ldots, x_m)$, where $F_i: \mathbb{R}^{n_i} \rightarrow \mathbb{R}$ is a continuously differentiable function. A set of strategies $x^* = (x^*_1, \ldots, x^*_m) \in \prod_{i=1}^M X_i$ is said to be in equilibrium if no player can reduce the incurred penalty (or increase the incurred reward) by unilaterally deviating from the chosen strategy. If each $F_i$ is convex on the set $X_i$, then the set of strategies $x^*$ is in equilibrium if and only if $\langle (x_i - x^*_i), \nabla_i F_i(x^*_i) \rangle \geq 0$. In other words, $x^*$ needs to be a solution of the VI $\langle (x - x^*), f(x^*) \rangle \geq 0$, where $f(x) = (\nabla F_1(x), \ldots, \nabla F_m(x))$. Nash games are closely related to {\em saddle point} problems \cite{mirror-prox,mirror-prox2,bo-sm:nips2012}. where we are given a function $F: X \times Y \rightarrow \mathbb{R}$, and the objective is to find a solution $(x^*, y^*) \in X \times Y$ such that
\[ F(x^*, y) \leq F(x^*, y^*) \leq F(x, y^*), \ \ \forall x \in X, \ \forall y \in Y \]
Here, $F$ is convex in $x$ for each fixed $y$, and concave in $y$ for each fixed $x$. Many equilibria problems in economics can be modeled using VIs \cite{nagurney:vibook}.
\section{Algorithms for Variational Inequalities}
\label{viola}
We briefly describe two algorithms for solving variational inequalities below: the projection method and the extragradient method. We conclude with a brief discussion of how these relate to reinforcement learning.
\subsection{Projection-Based Algorithms for VIs}
\label{egvi}
The basic projection-based method (Algorithm 1) for solving VIs is based on Theorem~\ref{projthm} introduced earlier.
\begin{algorithm}
\caption{The Basic Projection Algorithm for solving VIs.}
{\bf INPUT:} Given VI(F,K), and a symmetric positive definite matrix $D$.
\begin{algorithmic}[1]
\STATE Set $k=0$ and $x_k \in K$.
\REPEAT
\STATE Set $x_{k+1} \leftarrow \Pi_{K,D} (x_k - D^{-1} F(x_k))$.
\STATE Set $k \leftarrow k+1$.
\UNTIL{$x_k = \Pi_{K,D} (x_k - D^{-1} F(x_k))$}.
\STATE Return $x_k$
\end{algorithmic}
\end{algorithm}
Here, $\Pi_{K,D}$ is the projector onto convex set $K$ with respect to the natural norm induced by $D$, where $\| x \|^2_D = \langle x, D x \rangle$. It can be shown that the basic projection algorithm solves any $VI(F,K)$ for which the mapping $F$ is {\em strongly monotone} \footnote{ A mapping $F$ is {\em strongly monotone} if
$\langle F(x) - F(y), x - y \rangle \geq \mu \| x - y \|^2_2, \mu > 0, \forall x,y \in K$.} and {\em Lipschitz}.\footnote{A mapping $F$ is {\em Lipschitz} if $\| F(x) - F(y) \|_2 \leq L \|x - y \|_2, \forall x,y \in K$. }A simple strategy is to set $D = \alpha I$, where $\alpha > \frac{L^2}{2 \mu}$, and $L$ is the Lipschitz smoothness constant, and $\mu$ is the strong monotonicity constant. The basic projection-based algorithm has two critical limitations: it requires that the mapping $F$ be strongly monotone. If, for example, $F$ is the gradient map of a continuously differentiable function, strong monotonicity implies the function must be strongly convex. Second, setting the parameter $\alpha$ requires knowing the Lipschitz smoothness $L$ and the strong monotonicity parameter $\mu$. The extragradient method of Korpolevich \cite{korpelevich} addresses some of these concerns, and is defined as Algorithm 2 below.
\begin{algorithm}
\caption{The Extragradient Algorithm for solving VIs.}
{\bf INPUT:} Given VI(F,K), and a scalar $\alpha$.
\begin{algorithmic}[1]
\STATE Set $k=0$ and $x_k \in K$.
\REPEAT
\STATE Set $y_{k} \leftarrow \Pi_{K} (x_k - \alpha F(x_k))$.
\STATE \label{projstep} Set $x_{k+1} \leftarrow \Pi_{K} (x_k - \alpha F(y_k))$.
\STATE Set $k \leftarrow k+1$.
\UNTIL{$x_k = \Pi_{K} (x_k - \alpha F(x_k))$}.
\STATE Return $x_k$
\end{algorithmic}
\end{algorithm}
Figure~\ref{eg-vi-example} shows a simple example where Algorithm 1 fails to converge, but Algorithm 2 does. If the initial point $x_0$ is chosen to be on the boundary of $X$, using Algorithm 1, it stays on it and fails to converge to the solution of this VI (which is at the origin). If $x_0$ is chosen to be in the interior of $K$, Algorithm 1 will move towards the boundary. In contrast, using Algorithm 2, the solution can be found for any starting point. The extragradient algoriithm derives its name from the property that it requires an ``extra gradient" step (step 4 in Algorithm 2), unlike the basic projection algorithm given earlier as Algorithm 1. The principal advantage of the extragradient method is that it can be shown to converge under a considerably weaker condition on the mapping $F$, which now has to be merely monotonic: $\langle F(x) - F(y), x - y \rangle \geq 0$. The earlier Lipschitz condition is still necessary for convergence.
\begin{figure}[h]
\begin{center}
\begin{minipage}[t]{0.4\textwidth}
\includegraphics[width=\textwidth,height=1.5in]{eg-vi-problem.pdf}
\end{minipage}
\begin{minipage}[t]{0.4\textwidth}
\includegraphics[width=\textwidth,height=1in]{eg-one-step.pdf}
\end{minipage}
\end{center}
\caption{Left: This figure illustrates a VI where the basic projection algorithm (Algorithm 1) fails, but the extragradient algorithm (Algorithm 2) succeeds \cite{bertsekas:pdc}. Right: One iteration of the extradient algorithm.}
\label{eg-vi-example}
\end{figure}
The extragradient algorithm has been the topic of much attention in optimization since it was proposed, e.g., see \cite{iusem,khobotov,marcotte,peng:extragradient,nesterov:dualeg,solodov-svaiter}. Khobotov \cite{khobotov} proved that the extragradient method converges under the weaker requirement of {\em pseudo-monotone} mappings, \footnote{A mapping $F$ is {\em pseudo-monotone} if $\langle F(y), x - y \rangle \geq 0 \Rightarrow \langle F(x), x - y \rangle \geq 0, \ \forall x, y \in K$.} when the learning rate is automatically adjusted based on a local measure of the Lipschitz constant. Iusem \cite{iusem} proposed a variant whereby the current iterate was projected onto a hyperplane separating the current iterate from the final solution, and subsequently projected from the hyperplane onto the feasible set. Solodov and Svaiter \cite{solodov-svaiter} proposed another hyperplane method, whereby the current iterate is projected onto the intersection of the hyperplane and the feasible set. Finally, the extragradient method was generalized to the non-Euclidean case by combining it with the mirror-descent method \cite{nemirovski-yudin:book}, resulting in the so-called ``mirrror-prox" algorithm \cite{mirror-prox}.
\subsection{Variational Inequaities and Reinforcement Learning}
Variational inequalities also provide a useful framework for reinforcement learning \cite{ndp:book,sutton-barto:book}. In this case, it can be shown that the mapping $F$ for the VI defined by reinforcement learning is {\em affine} and represents a (linear) complementarity problem. For this case, a number of special properties can be exploited in designing a faster more scalable class of algorithms.
Recall from Theorem~\ref{projthm} that each VI(F,K) corresponds to solving a particular fixed point problem $x^* = \Pi_K(x^* - \gamma F(x^*))$, which led to the projection algorithm (Algorithm 1). Generalizing this, consider solving for the fixed point of a {\em projected equation} $x^* = \Pi_{\hat{S}} T(x^*)$ \cite{bertsekas:vi-tr09,gordon:lcp} for a functional mapping $T: \mathbb{R}^n \rightarrow \mathbb{R}^n$, where $\Pi_{\hat{S}}$ is the projector onto a low-dimensional convex subspace $\hat{S}$ w.r.t. some positive definite matrix $\Xi$, so that
\begin{equation}
\hat{S} = \{\Phi r | r \in \hat{R} \}, \ \ \hat{R} = \{r | \Phi r \in \hat{S} \} \Rightarrow \Phi r^* = \Pi_{\hat{S}} T(\Phi r^*)
\end{equation}
Here. $\Phi$ is an $n \times s$ matrix where $s \ll n$, and the goal is to make the computation depend on $s$, not $n$. Note that $x^* = \Pi_{\hat{S}} T(x^*)$ if and only if
\[ \langle (x^* - T(x^*), \Xi (x - x^*) \rangle \geq 0. \ \forall x \in \hat{S} \]
Following \cite{bertsekas:vi-tr09}, note that this is a variational inequality of the form $\langle F(x^*), (x - x^*) \rangle \geq 0$ if we identify $F(x) = \Xi (x - T(x))$, and in the lower-dimensional space, $\langle F(\Phi r^*), \Phi(r - r^*) \rangle, \ \forall r \in \hat{R}$. Hence, the projection algorithm takes on the form:
\[ x_{k+1} = \Pi_{\hat{S}} (x_k - \gamma D^{-1} \langle \Phi, F(\Phi x_k)) \rangle \]
It is shown in \cite{bertsekas:vi-tr09} that if $T$ is a contraction mapping, then $F(x) = \Xi (x - T(x))$ is strongly monotone. Hence, the above projection algorithm will converge to the solution $x^*$ of the VI for any given starting point $x_0 \in \hat{S}$. Now, the only problem is how to ensure the computation depends only on the size of the projected lower-dimensional space (i.e., $s$, not $n$). To achieve this, let us assume that the mapping $T(x) = Ax + b$ is affine, and that the constraint region $\hat{R}$ is polyhedral. In this case, we can use the following identities:
\[ \langle \Phi, F(\Phi, x) \rangle = \Phi^T \Xi F(\Phi x) = Cr - d, \ \ C = \Phi^T \Xi (I - A) \Phi, \ \ d = \Phi^T \Xi b \]
and the projection algorithm for this affine case can be written as:
\[ x_{k+1} = \Pi_{\hat{S}} (x_k - \gamma D^{-1} (C r - d) ) \]
One way to solve this iteratively is to compute a progressively more accurate approximation $C_k \rightarrow C$ and $d_k \rightarrow d$ by sampling from the rows and columns of the matrix $A$ and vector $b$, as follows:
\[ C_k = \frac{1}{k+1} \sum_{t=0}^k \phi(i_t) (\phi(i_t) - \frac{a_{i_t j_t}}{p_{i_t j_t}} \phi(j_t))^T, \ \ d_k = \frac{1}{k+1} \sum_{t=0}^k \phi(i_t) b_{i_t} \]
where the row sampling generates the indices $(i_0, i_1, \ldots)$ and the column sampling generates the transitions $((i_0, j_0), (i_1, j_1), \ldots, )$ in such a way that the relative frequency of row index $i$ matches the diagonal element $\xi_i$ of the positive definite matrix $\Xi$. Given $C_k$ and $d_k$, the solution can be found by $x^* \approx C_k^{-1} d_k$, or by using an incremental method. Computation now only depends on the dimension $s$ of the lower-dimensional space, not on the original high-dimensional space. Gordon \cite{gordon:lcp} proposes an alternative approach separating the projection of the current iterate on the low-dimensional subspace spanned by $\Phi$ from its projection onto the feasible set. Both of these approaches \cite{bertsekas:vi-tr09,gordon:lcp} have been only studied with the simple projection method (Algorithm 1), and can be generalized to a more powerful class of VI methods that we are currently developing.
\section*{Acknowledgements}
We like to acknowledge the useful feedback of past and present members of the Autonomous Learning Laboratory at the University of Massachusetts, Amherst. Principal funding for this research was provided by the National Science Foundation under the grant NSF IIS-1216467. Past work by the first author has been funded by NSF grants IIS-0534999 and IIS-0803288.
\chapter{Appendix: Technical Proofs}
\section{Convergence Analysis of Saddle Point Temporal Difference Learning}
\subsection*{Proof of Proposition 1}
We give a descriptive proof here. We first present the monotone operator corresponding to the bilinear saddle-point problem and then extend it to stochastic approximation case with certain restrictive assumptions, and use the result in \cite{sra2011optimization}.
The monotone operator $\Phi (x,y)$ with saddle-point problem $SadVal = {\inf _{x \in X}}{\sup _{y \in Y}}\phi (x,y)$ is a point-to-set operator
\[\Phi (x,y) = \{ {\partial _x}\phi (x,y)\} \times \{ - {\partial _y}\phi (x,y)\} \]
Where ${\partial _x}\phi (x,y)$ is the subgradient of $\phi(x,\cdot)$ over $x$ and ${\partial _y}\phi (x,y)$ is the subgradient of $\phi(\cdot,y)$ over $y$. For the bilinear problem in Equation (\ref{eq:minimax}), the corresponding $\Phi(x,y)$ is
\begin{equation}\label{eq:bilinearphi}
\Phi (x,y) = ({A^T}y,b - Ax)
\end{equation}
Now we verify that the problem (\ref{eq:lossfunc}) can be reduced to a standard bilinear minimax problem. To prove this we only need to prove $X$ in our RL problem is indeed a closed compact convex set. This is easy to verify as we can simply define $X = \{ x|{\left\| x \right\|_2} \le R\} $ where $R$ is large enough. In fact, the sparse regularization $h(x)= \rho ||x||_1$ helps $x_t$ stay within this $l_2$ ball.
Now we extend to the stochastic approximation case wherein the objective function $f(x)$ is given by the stochastic oracle, and in our case, it is $Ax-b$, where $A, b$ are defined in Equation (\ref{eq:UNIFY}), with Assumption 3 and further assuming that the noise $\varepsilon_t$ for $t$-th sample is i.i.d noise, then with the result in \cite{juditsky2008solving}, we can prove that the RO-TD algorithm converges to the global minimizer of
\[{x^*} = \arg {\min _{x \in X}}{\left\| {Ax - b} \right\|_m} + \rho {\left\| x \right\|_1}\]
Then we prove the error level of approximate saddle-point ${\bar x_t},{\bar y_t}$ defined in (\ref{eq:averaging}) is $\alpha_{t} L^2$. With the subgradient boundedness assumption and using the result in Proposition 1 in \cite{nedic2009subgradient}, this can be proved.
\section{Convergence Analysis of True Gradient Temporal Difference Learning}
We first present the assumptions for the MDP and basis functions,
which are similar to \cite{FastGradient:2009,ROTD:NIPS2012}.
\textbf{Assumption 1} (\textbf{MDP}): The underlying Markov Reward
Process (MRP) $M=(S,P,R,\gamma)$ is finite and mixing, with stationary
distribution $\pi$. The training sequence $({s_{t}},{a_{t}},s_{t}^{'})$
is an i.i.d sequence.
\textbf{Assumption 2} (\textbf{Basis Function}): The inverses $\mathbb{E}{[\phi_{t}{\phi_{t}^{T}}]^{-1}}$
and ${[{\phi_{t}}({\phi_{t}}-\gamma{\phi_{t}}^{'T})]^{-1}}$ exist.
This implies that $\Phi$ is a full column rank matrix. Also, assume
the features $(\phi_{t},\phi_{t}^{'})$ have uniformly bounded second
moments, and ${\left\Vert {\phi_{t}}\right\Vert _{\infty}}<+\infty,{\left\Vert {\phi{'_{t}}}\right\Vert _{\infty}}<+\infty$.
\\
Next we present the assumptions for the stochastic saddle point problem
formulation, which are similar to \cite{chen2013optimal,juditsky2008solving}.
\textbf{Assumption 3} (\textbf{Compactness}): Assume for the primal-dual
loss function,
\begin{equation}
\mathop{\min}\limits _{\theta\in X}\mathop{\max}\limits _{y\in Y}\left({L(\theta,y)=\left\langle {K(\theta),y}\right\rangle -{F^{*}}(y)}+h(\theta)\right),
\label{eq:primal-dual2}
\end{equation}
the sets $X,Y$ are closed compact sets.
\textbf{Assumption 4} ($F^{*}(\cdot)$): We assume that $F^{*}(\cdot)$
is a smooth convex function with Lipschitz continuous gradient, i.e.,
$\exists L_{F^{*}}$ such that $\forall x,y\in X$
\begin{equation}
{F^{*}}(y)-{F^{*}}(x)-\left\langle {\nabla{F^{*}}(x),y-x}\right\rangle \le\frac{{L_{{F^{*}}}}}{2}||y-x|{|^{2}}
\end{equation}
\textbf{Assumption 5} ($K(\theta)$): $K(\theta)$ is a linear mapping
which can be extended to Lipschitz continuous vector-valued mapping
defined on a closed convex cone $C_{K}$. Assuming $K(\theta)$ to
be $C_{K}$-convex, i.e., $\forall\theta,\theta'\in X,\lambda\in[0,1],$
\[
K(\lambda\theta+(1-\lambda)\theta'){\le_{{C_{K}}}}\lambda K(\theta)+(1-\lambda)K(\theta'),
\]
where $a{\le_{{C_{K}}}}b$ means that $b-a\in{C_{K}}$.
\textbf{Assumption 6} (\textbf{Stochastic Gradient}): In the stochastic saddle
point problem, we assume that there exists a stochastic oracle $\mathcal{SO}$
that is able to provide unbiased estimation with bounded variance
such that
\begin{equation}
\begin{array}{l}
\mathbb{E}[{{\cal F}^{*}}({y_{t}})]={\nabla{F^{*}}({y_{t}})}\\
\mathbb{E}[||{{\cal F}^{*}}({y_{t}})-\nabla{F^{*}}({y_{t}})||^2] \le{\sigma^2_{{F^{*}}}}
\end{array}
\end{equation}
\begin{equation}
\begin{array}{l}
\mathbb{E}[{\cal K_{\theta}}(\theta_{t})]=K({\theta_{t}})\\
\mathbb{E}[||{\cal K_{\theta}}(\theta_{t})-K({\theta_{t}})|{|^{2}}]\le\sigma_{K,\theta}^{2}
\end{array}
\end{equation}
\begin{equation}
\begin{array}{l}
\mathbb{E}[{{\cal K}_{y}(\theta_{t})^{T}}{y_{t}}]=\nabla K{(\theta_{t})^{T}}{y_{t}}\\
\mathbb{E}[||{{\cal K}_{y}(\theta_{t})^{T}}{y_{t}}-\nabla K{(\theta_{t})^{T}}{y_{t}}|{|^{2}}]\le\sigma_{K,y}^{2}
\end{array}
\end{equation}
where $\sigma_{{F^{*}}}$ , $\sigma_{K,\theta}$ and $\sigma_{K,y}$
are non-negative constants.
We further define
\begin{equation}
\sigma=\sqrt{\sigma_{{F^{*}}}^{2}+\sigma_{K,\theta}^{2}}+{\sigma_{K,y}}\label{eq:sigma}
\end{equation}
\subsection{Convergence Rate}
Here we discuss the convergence rate of the proposed algorithms. First let us review the nonlinear primal form
\begin{equation}
\mathop{\min}\limits _{\theta\in X}\left(\Psi(\theta)={F(K(\theta))+h(\theta)}\right)\label{eq:primal-1}
\end{equation}
The corresponding primal-dual formulation \cite{BOOK2011PROXSPLIT,PROXSPLITTING2011,POCK2011SADDLE}
of Equation (\ref{eq:primal-1}) is Equation (\ref{eq:primal-dual}).
Thus we have the general update rule as
\begin{equation}
\begin{array}{l}
{y_{t + 1}} = {y_t} + {\alpha _t}{{\cal K}_\theta }({\theta _t}) - {\alpha _t}{{\cal F}^*}({y_t})
{\rm {,\;}}
{\theta _{t + 1}} = {\rm{pro}}{{\rm{x}}_{{\alpha _t}h}}({\theta _t} - {\alpha _t}{{\cal K}_y}{({\theta _t})^T}{y_t})
\end{array}
\end{equation}
\textbf{Lemma 1 (Optimal Convergence Rate):}
The optimal convergence rate of \ref{eq:primal-dual} is
$O(\frac{{L_{{F^{*}}}}}{{N^{2}}}+\frac{{L_{K}}}{N}+\frac{\sigma}{{\sqrt{N}}})$.
\textbf{Proof:}
Equation (\ref{eq:primal-1}) can be easily converted to the
following primal-dual formulation
\begin{equation}
\mathop{\min}\limits _{y\in Y}\mathop{\max}\limits _{\theta\in X}\left({\left\langle {-K(\theta),y}\right\rangle +{F^{*}}(y)-h(\theta)}\right)\label{eq:primal-dual-2}
\end{equation}
Using the bounds proved in \cite{nesterov2004introductory,juditsky2008solving,chen2013optimal},
the optimal convergence rate of stochastic saddle-point problem is
$O(\frac{{L_{{F^{*}}}}}{{N^{2}}}+\frac{{L_{K}}}{N}+\frac{\sigma}{{\sqrt{N}}})$.
The GTD/GTD2 algorithms can be considered as using Polyak's algorithm
without the primal average step. Hence, by adding the primal average
step, GTD/GTD2 algorithms will become standard Polyak's algorithms
\cite{polyak1992acceleration}, and thus the convergence rates are
$O(\frac{{{L_{{F^{*}}}}+{L_{K}}+\sigma}}{{\sqrt{N}}})$ according
to \cite{RobustSA:2009}. So we have the following propositions.
\textbf{Proposition 1} The convergence rates of GTD/GTD2 algorithms
with primal average are $O(\frac{{{L_{{F^{*}}}}+{L_{K}}+\sigma}}{{\sqrt{N}}})$,
where ${L_{K}}=||{\Phi^{T}}\Xi(\Phi-\gamma{\Phi^{'T}})|{|^{2}}$,
for GTD, ${L_{{F^{*}}}}=1$ and for GTD2, ${L_{{F^{*}}}}=||M|{|_{2}}$.
Now we consider the acceleration using the SMP algorithm, which incorporates
the extragradient term. According to \cite{juditsky2008solving} which extends the SMP algorithm to solving saddle-point problems and variational inequality
problems, the convergence rate is accelerated to $O(\frac{{{L_{{F^{*}}}}+{L_{K}}}}{N}+\frac{\sigma}{{\sqrt{N}}})$. Consequently,
\textbf{Proposition 2 }The convergence rate of the GTD2-MP algorithm is
$O(\frac{{{L_{{F^{*}}}}+{L_{K}}}}{N}+\frac{\sigma}{{\sqrt{N}}})$.
\subsection{Value Approximation Error Bound}
\label{eq:mspbe-2}
One key question is how to give the error bound of $||V-{V_{\theta}}||$
given that of $||K(\theta)||$. Here we use the result in \cite{DantzigRL:2012},
which is similar to the one in \cite{Yu08newerror}.
\textbf{Lemma 2 \cite{DantzigRL:2012}:} For any ${V_{\theta}}=\Phi\theta$,
the following component-wise equality holds
\begin{equation}
V-{V_{\theta}}={(I-\gamma{\Pi^{\Xi}}P)^{-1}}\left({\left({V-{\Pi^{\Xi}}V}\right)+\Phi{({\Phi^{T}}\Xi\Phi)^{-1}}K(\theta)}\right)\label{eq:ds_v_ax-b-1}
\end{equation}
\textbf{Proof}:
Use the equality $V=TV$ and ${V_{\theta}}={\Pi^{\Xi}}{V_{\theta}}$,
where the first equality is the Bellman equation, and the second is that
$V_{\theta}$ lies within the spanning space of $\Phi$.
we have
\begin{equation}
\begin{array}{l}
V-{\Pi^{\Xi}}V\\
=V-{\Pi^{\Xi}}TV+({V_{\theta}}-{\Pi^{\Xi}}T{V_{\theta}})-({V_{\theta}}-{\Pi^{\Xi}}T{V_{\theta}})\\
=(I-\gamma{\Pi^{\Xi}}P)(V-{V_{\theta}})+{\Pi^{\Xi}}({V_{\theta}}-T{V_{\theta}})
\end{array}
\end{equation}
After rearranging the equation, we have Equation (\ref{eq:ds_v_ax-b-1}) and find that
\begin{equation}
{\Pi^{\Xi}}({V_{\theta}}-T{V_{\theta}})=-\Phi{({\Phi^{T}}\Xi\Phi)^{-1}}K(\theta)
\end{equation}
\textbf{Proposition }3: For GTD/GTD2, the prediction error of $||V-{V_{\theta}}||$
is bounded by $||V-{V_{\theta}}|{|_{\infty}}\le\frac{{L_{\phi}^{\Xi}}}{{1-\gamma}}\cdot O\left({\frac{{{L_{{F^{*}}}}+{L_{K}}+\sigma}}{{\sqrt{N}}}}\right)$
; For GTD2-MP, it is bounded by $||V-{V_{\theta}}|{|_{\infty}}\le\frac{{L_{\phi}^{\Xi}}}{{1-\gamma}}\cdot O\left({\frac{{{L_{{F^{*}}}}+{L_{K}}}}{N}+\frac{\sigma}{{\sqrt{N}}}}\right)$,
where $L_{\phi}^{\Xi}={\max_{s}}||{({\Phi^{T}}\Xi\Phi)^{-1}}\phi(s)|{|_{1}}$.
\textbf{Proof:}
From Lemma 2, we have
\begin{equation}
||V-{V_{\theta}}|{|_{\infty}}\le||{(I-\gamma{\Pi^{\Xi}}P)^{-1}}|{|_{\infty}}\cdot\left({||V-{\Pi^{\Xi}V}|{|_{\infty}}+L_{\phi}^{\Xi}||K(\theta)|{|_{\infty}}}\right)
\end{equation}
Using the results in Proposition 1 and Proposition 2, we have for GTD
and GTD2,
\begin{equation}
||V-{V_{\theta}}|{|_{\infty}}\le||{(I-\gamma{\Pi^{\Xi}}P)^{-1}}|{|_{\infty}}\cdot\left({||V-{\Pi^{\Xi}V}|{|_{\infty}}+L_{\phi}^{\Xi}\cdot O\left({\frac{{{L_{{F^{*}}}}+{L_{K}}+\sigma}}{{\sqrt{N}(1-\gamma)}}}\right)}\right)
\end{equation}
For GTD2-MP,
\begin{equation}
||V-{V_{\theta}}|{|_{\infty}}\le||{(I-\gamma{\Pi^{\Xi}}P)^{-1}}|{|_{\infty}}\cdot\left({||V-{\Pi^{\Xi}V}|{|_{\infty}}+L_{\phi}^{\Xi}\cdot O(\frac{{{L_{{F^{*}}}}+{L_{K}}}}{N}+\frac{\sigma}{{\sqrt{N}}})}\right)
\end{equation}
If we further assume a rich expressive hypothesis
space $\mathcal{H}$, i.e., ${\Pi^{\Xi}}P=P,{\Pi^{\Xi}}R=R$, $||V-{\Pi^{\Xi}}V|{|_{\infty}}=0,||{(I-\gamma{\Pi^{\Xi}}P)^{-1}}|{|_{\infty}}=\frac{1}{{1-\gamma}}$,
then for GTD and GTD2, we have
\begin{equation}
||V-{V_{\theta}}|{|_{\infty}}\le\frac{{L_{\phi}^{\Xi}}}{{1-\gamma}}\cdot O\left({\frac{{{L_{{F^{*}}}}+{L_{K}}+\sigma}}{{\sqrt{N}}}}\right)
\end{equation}
For GTD2-MP, we have
\begin{equation}
||V-{V_{\theta}}|{|_{\infty}}\le\frac{{L_{\phi}^{\Xi}}}{{1-\gamma}}\cdot O\left({\frac{{{L_{{F^{*}}}}+{L_{K}}}}{N}+\frac{\sigma}{{\sqrt{N}}}}\right)
\end{equation}
\bibliographystyle{unsrtnat}
|
\section{Introduction}
\subsection{Motivation and definitions}
Circulant graphs, or multi-loop networks as used in computer science literature, are basic structures for interconnection networks \cite{Bermond1995}. As such a lot of research on circulant graphs has been done in more than three decades, leading to a number of results on various aspects of circulant graphs \cite{Bermond1995, Fertin2004,Gauyacq1998,Gomez2007, Hwang2001, Mans2004, Stojmenovic1997,Thomson2008,Thomson2010,Thomson2012}. Nevertheless, our knowledge on how circulant networks behave with regard to information dissemination is very limited. For example, our understanding to some basic communication-related invariants for circulant graphs such as the arc-forwarding, edge-forwarding and optical indices is quite limited. The purpose of this paper is to study these invariants with a focus on circulant networks of degree 4.
Given an integer $n \ge 3$, denote by $\ZZ_n$ the group of integers modulo $n$ with operation the usual addition. Given $S \subset \ZZ_n$ such that $ 0 \not \in S$ and $s\in S$ implies $-s\in S$, the \textit{circulant graph} $C_n(S)$ of order $n$ with respect to $S$ is defined to have vertex set $\ZZ_n$ such that $i, j \in \ZZ_n$ are adjacent if and only if $i-j\in S$. (In other words, a circulant graph is a Cayley graph on $\ZZ_n$.) In the case when $S = \{a, b, n-a, n-b\}$, where $a, b, n-a, n-b$ are pairwise distinct elements of $\ZZ_n$, $C_n(S)$ is a 4-regular graph (that is, every vertex has degree 4) and we use $C_n(a, b)$ in place of $C_n(S)$. In this paper we deal with circulant graphs $C_n(1, s)$ for some $s \in \ZZ_n \setminus \{-1, 0, 1, n/2\}$. (Note that when $n$ and $a$ are coprime, $C_n(a, b)$ is isomorphic to $C_n(1,s)$, where $s \equiv a^{-1}b\mod n$). Without loss of generality, we assume $1<s<n/2$. Just like any other Cayley graph, $C_n(1, s)$ is vertex-transitive, that is, for any $i, j \in \ZZ_n$ there exists a permutation of $\ZZ_n$ that preserves the adjacency relation of $C_n(1, s)$ and maps $i$ to $j$. (In fact, for fixed $i, j$ this permutation can be chosen as $x \mapsto x+(j-i)$, $x \in \ZZ_n$ with operation undertaken in $\ZZ_n$.)
A network can be represented by an undirected graph $G=(V(G), E(G))$, where the node set $V(G)$ represents the set of processors or routers, and the edge set $E(G)$ represents the set of physical links. So we will use the words `graph' and `network' interchangeably. We assume the full duplex model, that is, an edge is regarded as two \textit{arcs} with opposite directions over which messages can be transmitted concurrently. A connection request (or a \textit{request} for short) is an ordered pair of distinct nodes $(x,y)$ for which a path $P_{x, y}$ with orientation from $x$ to $y$ in $G$ must be set up to transmit messages from $x$ to $y$. In this paper we only consider all-to-all communication, or equivalently, the \textit{all-to-all request set} for which one path from every node to every other node must be set up in order to fulfil communications. (In the literature other types of request sets have also been studied.) We call a set of paths $R = \{P_{x, y}: x, y \in V(G), x \not=y\}$ an \textit{all-to-all routing} (or a \textit{routing} for short) in $G$, where $P_{x, y}$ is not necessarily the same as $P_{y, x}$. The load of an edge $e$ of $G$ with respect to $R$, denoted by $\pi(G,R,e)$, is the number of paths in $R$ passing through $e$ in either directions. Similarly, the load of an arc $a$ of $G$ with respect to $R$, denoted by ${\overrightarrow{\pi}}(G,R,a)$, is the number of paths in $R$ passing through $a$ along its direction. Define
\begin{equation}
\label{eq:load}
\pi(G, R):=\max_{e \in E(G)} \pi(G,R,e),\quad {\overrightarrow{\pi}}(G, R):=\max_{a\in A(G)} {\overrightarrow{\pi}}(G, R, a),
\end{equation}
where $A(G)$ is the set of arcs of $G$. Define
\begin{equation}
\label{eq:pidef}
\pi(G):=\min_{R}\pi(G,R),\quad {\overrightarrow{\pi}}(G):=\min_{R}{\overrightarrow{\pi}}(G,R)
\end{equation}
and call them the \emph{edge-forwarding} and \emph{arc-forwarding indices} of $G$ \cite{Beauquier1997,Heydemann1989}, respectively, where the minimum is taken over all routings $R$ for $G$. Obviously, we have
\begin{equation}
\label{eq:pi}
{\overrightarrow{\pi}}(G)\geq\pi(G)/2.
\end{equation}
The \textit{edge-forwarding index problem} is the one of finding $\pi(G)$ for a given graph $G$, and the \textit{arc-forwarding index problem} is understood similarly.
In practical terms, the edge-forwarding and arc-forwarding indices measure the minimum heaviest load on edges and arcs of a given network, respectively, with respect to all-to-all communication. If the network is all-optical, another important problem is to minimise the number of wavelengths used such that any two paths having an edge (or arc) in common are assigned distinct wavelengths. Regarding wavelengths as colours, these problems can be formulated as the following {\em path colouring problems}. Given a routing $R$ for $G$, an assignment of one colour to each path in $R$ is called an {\em edge-conflict-free colouring} of $R$ if any two paths having an edge in common (regardless of the orientation of the paths) receive distinct colours, and an {\em arc-conflict-free colouring} of $R$ if any two paths having an arc in common (with the same orientation as the paths) receive distinct colours. (An edge-conflict-free colouring is called \emph{valid} in \cite{Gargano2000}.) Define $w(G, R)$ (${\overrightarrow{w}}(G,R)$, respectively) to be the minimum number of colours required in an edge-conflict-free (arc-conflict-free, respectively) colouring of $R$. Define
\begin{equation}
\label{eq:w}
w(G):=\min_{R} w(G,R),\quad {\overrightarrow{w}}(G):=\min_{R}{\overrightarrow{w}}(G,R)
\end{equation}
and call them the {\em undirected} and {\em directed optical indices} of $G$, respectively,
where the minimum is taken over all routings $R$ for $G$.
Since the number of colours needed is no less than the number of paths on a most loaded edge (or arc in the directed version), we have (see e.g.~\cite{Bermond2000})
\begin{equation}
\label{eq:piw}
w(G)\ge\pi(G),\quad {\overrightarrow{w}}(G)\ge{\overrightarrow{\pi}}(G).
\end{equation}
In general, equality in (\ref{eq:piw}) is not necessarily true (see e.g.~\cite{Kosowski2009,Yuan2007}). The \textit{routing and wavelength assignment problem} is the problem of computing $w(G)$, and its \textit{oriented version} is the one of finding ${\overrightarrow{w}}(G)$.
\subsection{Literature review}
The study of the forwarding indices has been intensive in the literature.
Heydemann et al. \cite{Heydemann1989} proposed the edge-forwarding index problem and obtained basic results on this invariant, including upper bounds for the Cartesian product of graphs. In \cite{Sole1994} it was proved that orbital regular graphs (which are essentially Frobenius graphs \cite{Fang1998} except cycles and stars) achieve the smallest possible edge-forwarding index. In \cite{Thomson2008,Thomson2010,Thomson2012}, Thomson and Zhou gave formulas for the edge-forwarding and arc-forwarding indices of two interesting families of Frobenius circulant graphs. The exact value of edge-forwarding index of some other graphs have also been computed, including Kn\"odel graphs \cite{Fertin2004} and recursive circulant graphs \cite{Gauyacq1998}. However, in general it is difficult to find the exact value or a good estimate of the edge-forwarding or arc-forwarding index of a graph, even for some innocent-looking classes of graphs such as circulant graphs. The authors of \cite{Xu2007} obtained lower and upper bounds on the edge-forwarding index of a general circulant graph. However, these bounds are difficult to compute in general. Also, a uniform routing of shortest paths may not exist for circulant graphs, just as the case for Cayley graphs in general \cite{Heydemann1997}.
The reader is referred to the recent survey \cite{Xu2013} for the state-of-the-art on edge-forwarding and arc-forwarding indices of graphs.
The routing and wavelength assignment problem has also received considerable attention due to the importance of optical networking.
The authors of \cite{Beauquier1997} surveyed theoretical results and asked a few questions on the routing and wavelength assignment problem for a general request set as well as the all-to-all request set, especially for trees, rings, tori, meshes and hypercubes.
In \cite{Gargano2000} a survey of results on several versions of the routing and wavelength assignment problem was given and the exact value of the directed optical index of stars was obtained.
The exact value of the directed optical index was computed for trees \cite{Gargano1997}, rings \cite{Bermond2000}, trees of rings \cite{Beauquier:1999b}, hypercubes \cite{Bermond2000}, a few families of recursive circulant graphs \cite{Amar2001}, Cartesian sum of complete graphs \cite{Beauquier1999}, Cartesian product of paths with equal even lengths \cite{Beauquier1999}, and Cartesian product of some graphs such as rings \cite{Beauquier1999, Schroder1997}. While the exact value of the undirected optical index has been obtained for rings and hypercubes \cite{Bermond2000}, determining this invariant remains open for many families of graphs \cite{Beauquier1997}. For the undirected routing and wavelength assignment problem for arbitrary request sets on trees of rings, the best known result is a 2.75-approximation algorithm \cite{Bian2009} and a 2-approximation algorithm \cite{Deng2003} for a subfamily of such graphs.
In \cite{Kosowski2009} it was proved that the problem of deciding whether $\pi(G)\le 3$ or $w(G) \le 3$ is NP-complete.
\subsection{Main results}
{\em
In what follows we assume that $n$ and $s$ are integers with $n \ge 5$ and $1 < s < n/2$, and $q$ and $r$ are integers defined by
$$
q := \lfloor n/s\rfloor,\;\, n := q s + r.
$$
}
Observe that $0 \le r < s$ and $q \ge 2$ as $s < n/2$.
Denote by $x\oplus y$ ($x \ominus y$, respectively) the integer $x+y\mod n$ ($x - y\mod n$, respectively) between $0$ and $n-1$, where $x,y\in \ZZ_n$. Denote
$$
\epsilon(x) := \left\{
\begin{array}{ll}
1,\;\; \mbox{if $x$ is an odd integer}\\[0.2cm]
0,\;\; \mbox{if $x$ is an even integer.}
\end{array}
\right.
$$
The first main result in this paper is as follows.
\begin{thm}
\label{thm:tr1}
The following hold:
\begin{itemize}
\item[\rm (a)]
if $2 \le s \le \sqrt{n} - 1$, then
\begin{equation}
\label{eq:tf1}
\frac{n^2 - \epsilon(n)}{8(s+1)} \leq \frac{\pi(C_n(1,s))}{2} \le {\overrightarrow{\pi}}(C_n(1,s)) \leq \frac{(n-r)( n+ r+2)+s^2}{8s};
\end{equation}
\item[\rm (b)]
if $ s = \sqrt{n}$, then
\begin{equation}
\label{eq:tf2}
\frac{\sqrt{n}\ (n - 1)}{8}
\leq
\frac{\pi(C_n(1,s))}{2} \le {\overrightarrow{\pi}}(C_n(1,s))
\leq
\frac{\sqrt{n}\ (n - \epsilon(s))}{8};
\end{equation}
\item[\rm (c)]
if $\sqrt{n} + 1 \le s < n/2$, then
\begin{equation}
\label{eq:tf3}
\max\left\{\frac{n^2 - \epsilon(n)}{8(s+1)},\frac{(n-1)(\sqrt{2n}-7)^3}{24n} \right\}
\leq
\frac{\pi(C_n(1,s))}{2} \le {\overrightarrow{\pi}}(C_n(1,s)) $$ $$ \hspace{3in}
\leq
\frac{s^2 (n + r+ 2) - \epsilon(s) (n-r)}{8s}.
\end{equation}
\end{itemize}
\end{thm}
Let $\mathtt{ratio}$ denote the ratio of the upper bound to the lower bound in the same equation above. In (\ref{eq:tf1}), $\mathtt{ratio} = \frac{s+1}{s} \left(\left(1-\frac{r}{n}\right)\left(1+\frac{r+2}{n}\right) + \frac{s^2}{n^2}\right)\frac{n^2}{n^2 - \epsilon(n)}$, which is asymptotically $(s+1)/s$ ($\le 3/2$) as $n \rightarrow \infty$.
In (\ref{eq:tf2}), $\mathtt{ratio} = \frac{n-\epsilon(n)}{n-1}$, which tends to 1 as $n \rightarrow \infty$.
By using the lower bound $(n^2-\epsilon(n))/8(s+1)$ and the upper bound $s(n+r+2)/8$ in (\ref{eq:tf3}), we have $\mathtt{ratio} \le (s^2+s)$$ (n+r+2) / (n^2 - \epsilon(n))$. If $ s = c\sqrt{n}$, for $c\in(1,\sqrt{3/2}]$, then $\mathtt{ratio} \le c^2 +o(1)$, which is at most $3/2$ asymptotically.
If $\sqrt{3n/2} < s < n/2$, then the lower bound in (\ref{eq:tf3}) is equal to $(n-1)(\sqrt{2n}-7)^3/24n$ and $\mathtt{ratio} \le 3s n (n+1)/(n-1)(\sqrt{2n}-7)^3$. In the latter case, the upper bound increases with $s$ and can be $O(\sqrt{n})$ in the worst case scenario when $s \approx cn$ for some constant $c < 1/2$.
We will prove the lower bounds in (\ref{eq:tf1})-(\ref{eq:tf3}) in the next section. The upper bounds will be proved in Section \ref{sec:ub}; see Lemma \ref{lem:ub1} which gives more information and better upper bounds in some cases. To establish the upper bounds, we will give a specific routing (see Construction \ref{const}), which can be viewed as an approximation algorithm for computing ${\overrightarrow{\pi}}(C_n(1,s)$ and $\pi(C_n(1,s))$. From the discussion above, for $2 \le s \le \sqrt{3n/2}$, the performance ratio of this algorithm is at most $3/2$ asymptotically. So we obtain the following corollary of Theorem \ref{thm:tr1}.
\begin{cor}
\label{cor:tr1}
There is a 1.5-factor approximation algorithm to solve the edge-forwarding and arc-forwarding problems for 4-regular circulant graphs $C_n(1, s)$ with $n$ sufficiently large and $2 \le s \le \sqrt{3n/2}$.
\end{cor}
In the worst case when $s \approx cn$ is large, where $c < 1/2$ is a constant, the ratio obtained from (\ref{eq:tf3}) (and that of the approximation algorithm from Construction \ref{const}) is $O(\sqrt{n})$. It seems that this large ratio is due to the fact that the lower bound in (\ref{eq:tf3}) is unsatisfactory when $s$ is large. Our next result shows that sometimes we can significantly improve this lower bound for large $s$. This enhanced lower bound together with the upper bound in (\ref{eq:tf3}) implies that for $s \approx c n$ our algorithm can achieve a constant performance ratio $1/c$ in some cases.
\begin{thm}
\label{thm:lwr3}
If $r\leq q$ or $r+q\ge s+1$, then
\begin{equation}
\label{eq:lwr3}
\pi(C_n(1,s))\geq \frac{1}{2} \left\lfloor\frac{(s+1)^2}{2}\right\rfloor.
\end{equation}
\end{thm}
We prove Theorem \ref{thm:lwr3} by computing the sum of the distances between all pairs of nodes in $C_n(1,s)$, which is done by investigating an equivalent problem \cite{Zerovnik1993} for the integer lattice $\ZZ^2$.
The second main result in this paper is the following theorem on the optical indices of $C_n(1,s)$. Denote
\begin{equation}
\label{eq:ka}
\kappa(a) := a+\frac{\epsilon(s)+\epsilon(q)}{2}.
\end{equation}
\begin{thm}
\label{thm:tr2}
The following hold:
\begin{itemize}
\item[\rm (a)] if $2 \le s \le \sqrt{n-r + (\kappa(-2))^2} + \kappa(-2)$, then
\begin{equation}
\label{eq:thw1}
\frac{n^2 - \epsilon(n)}{8(s+1)} \leq
{\overrightarrow{w}}(C_n(1,s)) \leq
\frac{s+2}{24}\left(6q^2 + 3 q (s+4) + s (4s+10) + \epsilon(q) (2q + 3s +3)\right) ;
\end{equation}
\item[\rm (b)] if $ \sqrt{n-r+(\kappa(-1))^2} +\kappa(-1) \le s \le \sqrt{n-r + (\kappa(0))^2} + \kappa(0)$, then
\begin{equation}
\label{eq:thw2}
\frac{n^2 - \epsilon(n)}{8(s+1)} \leq {\overrightarrow{w}}(C_n(1,s))\le
\frac{ q(q+2)( 5q+2)}{24} + \frac{s (s+2) (2s+5)}{6}+\epsilon(q)\frac{5 q^2 + 13 q + 7}{8};
\end{equation}
\item[\rm (c)] if $\sqrt{n-r+(\kappa(1))^2}+ \kappa(1) \le s < n/2$, then
\begin{equation}
\label{eq:thw3}
\max\left\{ \frac{n^2 - \epsilon(n)}{8(s+1)},\frac{(n-1)(\sqrt{2n}-7)^3}{24n} \right\}
\leq
{\overrightarrow{w}}(C_n(1,s)) \hspace{1in} $$ $$ \hspace{1in}\le
\frac{q(q+2) (q+10)}{24} + \frac{s(s+2)(q+1)}{2} +\epsilon(q)\frac{(q+5)^2 + 4 (s+1)^2 }{8}.
\end{equation}
\end{itemize}
Moreover, the same lower and upper bounds are valid if ${\overrightarrow{w}}(C_n(1,s))$ is replaced by $w(C_n(1,s))/2$ in (\ref{eq:thw1})-(\ref{eq:thw3}).
\end{thm}
Let $\mathsf{ratio}$ denote the ratio of the upper bound to the lower bound in the same equation above.
In (\ref{eq:thw1}), $\mathsf{ratio} \le 2(1+\frac{1}{s})(1+\frac{2}{s}) \left(1+ \frac{s^2}{ 2n} + \frac{2s^4}{3 n^2} \right)\frac{n^2}{n^2-\epsilon(n)}+O\left(\frac{1}{\sqrt{n}}\right)$, which is $2(1+\frac{1}{s})(1+\frac{2}{s})$ asymptotically when $s^2 = o(n)$. When $s^2 = \Omega(n)$, in (\ref{eq:thw1}) we have $\mathsf{ratio} \le 13/3$ asymptotically since $s \le \sqrt{n-r + (\kappa(-2))^2} + \kappa(-2)$.
In (\ref{eq:thw2}), $\mathsf{ratio} \le 13/3$ asymptotically as $s\approx q \approx \sqrt{n}$ in case (b). Let
\begin{equation}
\label{eq:delta}
\delta(n) := 3n(n^2 - \epsilon(n))/(n-1)(\sqrt{2n}-7)^3-1.
\end{equation}
Then $\delta(n) + 1 > 3\sqrt{n}/2\sqrt{2} > 1$ if $n \ge 25$, $\delta(n) \approx 3\sqrt{n}/2\sqrt{2}$ as $n \rightarrow \infty$, and $\delta(n) +1 \le 91\sqrt{n}/80$ for sufficiently large $n$.
If $s \le \delta(n)$, then the lower bound in (\ref{eq:thw3}) is equal to $(n^2 - \epsilon(n))/8(s+1)$ and $\mathtt{ratio}\le \frac{q (q+2)(q+10)(s+1)}{3 (n^2-\epsilon(n))} + \frac{4 (s+1) (s+2)(n-r+s)}{ n^2-\epsilon(n)} + O(\frac{1}{\sqrt{n}})$. This upper bound increases with $s$ and approaches a constant between $13/3$ and $4.85$, depending on the value of $s$, as $n \rightarrow \infty$. The upper bounds in (\ref{eq:thw1})-(\ref{eq:thw3}) will be proved by giving a specific colouring (see Construction \ref{colouring}) of the routing obtained from Construction \ref{const}, and this can be viewed as an approximation algorithm for the routing and wavelength assignment problem and its oriented version for $C_n(1, s)$. Thus the discussion above yields the following corollary of Theorem \ref{thm:tr2}.
\begin{cor}
\label{cor:tr2}
There is a $4.85$-factor approximation algorithm to solve the routing and wavelength assignment problem and its oriented version for 4-regular circulant graphs $C_n(1, s)$ with $n$ sufficiently large and $3 \le s \le 3\sqrt{n}/2\sqrt{2} - 1$.
\end{cor}
If $\delta(n) < s < n/2$, then the lower bound in (\ref{eq:thw3}) is equal to $(n-1)(\sqrt{2n}-7)^3/24n$ and the $\mathsf{ratio}$ from (\ref{eq:thw3}) (and so the approximation ratio of the algorithm from Construction \ref{colouring}) is at most $\frac{n} {(n - 1)}\left(\frac{ q(q+2) (q+10)}{(\sqrt{2n}-7 )^3} + \frac{s(s+2)(q+1)+\epsilon(q) s^2 }{2(\sqrt{2n}-7 )^3}\right)+ O(\frac{1}{\sqrt{n}})$. This upper bound increases with $s$ and is $O(\sqrt{n})$ in the worst case scenario when $s \approx cn$ for some constant $c < 1/2$. However, in the case when $r \le q$ or $r+q \ge s+1$, our stronger lower bound on $w(C_n(1,s))$ obtained from (\ref{eq:lwr3}) and (\ref{eq:piw}) implies that the approximation ratio is at most $4(1/c+ \epsilon(q))$ asymptotically.
The rest of the paper is organised as follows.
We will prove the lower bounds in the theorems above in the next section. In Section \ref{sec:ub} we will establish the upper bounds in Theorem \ref{thm:tr1} by devising a specific routing (Construction \ref{const}). In Section \ref{sec:wave} we will prove the upper bounds in Theorem \ref{thm:tr2} by giving a specific colouring (Construction \ref{colouring}) for the routing obtained from Construction \ref{const}. Note that the lower bounds in (\ref{eq:thw1})-(\ref{eq:thw3}) are obtained from (\ref{eq:piw}) and the lower bounds in Theorem \ref{thm:tr1}.
\section{Lower bounds}
\label{sec:lwrs}
\subsection{Two lower bounds}
\label{subsec:two}
Given a graph $G$ and $U \subset V(G)$, let $\delta(U)$ denote the set of edges of $G$ with one end in $U$ and the other end in $\overline{U} = V(G) \setminus U$. Let $R^*$ be a routing of $G$ such that $\pi(G)=\pi(G, R^*)$. Then $\pi(G)$ is the maximum load on an edge of $G$ under $R^*$. The total load on the edges of $\delta(U)$ under $R^*$ is thus at most $\pi(G)|\delta(U)|$. On the other hand, there are exactly $2 |U| |\overline{U}|$ paths in $R^*$ with one end in $U$ and the other end in $\overline{U}$, and each of them uses at least one edge of $\delta(U)$. Therefore,
\begin{equation}
\label{eq:is2}
\pi(G)|\delta(U)| \geq 2 |U| |\overline{U}|.
\end{equation}
\begin{lem}
\label{lem:lwr1}
$$
\pi(C_n(1,s)) \ge \frac{\lfloor n/2\rfloor\lceil n/2\rceil}{s+1} = \frac{n^2 - \epsilon(n)}{4(s+1)}.
$$
\end{lem}
\begin{proof}
We apply (\ref{eq:is2}) to $C_n(1,s)$. Choose $U=\{0,\dots,\lfloor n/2\rfloor-1\}\subset\mathbb{Z}_n$ so that $|U|=\lfloor n/2\rfloor$ and $|\overline{U}|=\lceil n/2\rceil $. Consider the neighbours $i+s$, $i-s$ of $i \in U$. We have: $\lfloor n/2\rfloor \le i+s \le n-1$ if and only if $\lfloor n/2\rfloor-s\le i\le \lfloor n/2\rfloor-1$, and $i - s$ ($\equiv n-(s-i) \mod n$) lies in $\overline{U}$ if and only if $0\le i\le s-1$. Thus $\delta(U)$ consists of edges $\{i, i+s\}$ ($\lfloor n/2\rfloor-s\le i\le \lfloor n/2\rfloor-1$), $\{i,i-s\}$ ($0\le i\le s-1$), $\{0,n-1\}$ and $\left\{\lfloor n/2\rfloor-1,\lfloor n/2\rfloor\right\}$. Hence $|\delta(U)|=2s+2$. This together with (\ref{eq:is2}) yields $\pi\big(C_n(1,s)\big) \geq \lfloor n/2\rfloor\lceil n/2\rceil/(s+1)$.
\end{proof}
As observed in \cite{Heydemann1989}, we have
\begin{equation}
\label{eq:trvl}
\pi(G)\geq\frac{\sum_{(x, y)\in V(G) \times V(G)}d(x, y)}{|E(G)|} = \frac{n(n-1)\overline{d}(G)}{|E(G)|},
\end{equation}
where $d(x, y)$ is the distance between $x$ and $y$ in $G$ and $\overline{d}(G)$ is the mean distance among all unordered pairs of vertices of $G$. It was proved in \cite[Theorem 4.6]{Hwang2003} that, if $G$ is a circulant network of order $n$ and degree $4$, then
$$
\overline{d}(G) \geq\frac{(\sqrt{2n}-7)^3}{6n}.
$$
Applying this and (\ref{eq:trvl}) to $C_n(1,s)$, we obtain:
\begin{lem}
\label{lem:lwr2}
$$
\pi(C_n(1,s))\geq \frac{(n-1)(\sqrt{2n}-7)^3}{12n}.
$$
\end{lem}
It can be verified that the lower bound in Lemma \ref{lem:lwr1} is no less than that in Lemma \ref{lem:lwr2} if and only if $s \le 3n(n^2 - \epsilon(n))/(n-1)(\sqrt{2n}-7)^3 -1$. Thus the lower bounds in (\ref{eq:tf1})-(\ref{eq:tf3}) follow from (\ref{eq:pi}) and Lemmas \ref{lem:lwr1} and \ref{lem:lwr2} immediately.
\subsection{Proof of Theorem \ref{thm:lwr3}}
\label{subsec:rq}
To prove Theorem \ref{thm:lwr3} we need two results from \cite{Zerovnik1993}.
Let $\ZZ^2$ be the $2$-dimensional $\ZZ$-module lattice. Define $l:\ZZ^2 \rightarrow \ZZ_n$ by
$$
l(\mathbf{x}):=x_1\oplus x_2 s,\;\, \mathbf{x}=(x_1,x_2) \in \ZZ^2.
$$
We may view $l$ as a labelling that labels each point $(x_1, x_2)$ of $\ZZ^2$ by the node $x_1 \oplus x_2 s$ of $C_n(1,s)$. We observe that when two points $(x_1,x_2)$ and $(y_1,y_2)$ are neighbours in the lattice $\ZZ^2$ (that is, either $x_1=y_1$ and $|x_2-y_2|=1$, or $|x_1-y_1|=1$ and $x_2=y_2$), the corresponding labels $l(x_1,x_2)$ and $l(y_1,y_2)$ are adjacent nodes of $C_n(1, s)$.
Denote by $\norm{.}$ the $L_1$-norm in $\ZZ^2$ defined by
$$
{\norm{\mathbf{x}}} := |x_1|+|x_2|,\;\, \mathbf{x}=(x_1,x_2) \in \ZZ^2.
$$
The {\em length} of a path $\mathbf{x}_0, \mathbf{x}_1,\dots,\mathbf{x}_k$ in $\ZZ^2$, connecting $\mathbf{x}_0$ and $\mathbf{x}_k$ is defined as $k$, and the {\em distance} between two points of $\ZZ^2$ is defined to be the length of a shortest path in $\ZZ^2$ connecting them, where $\mathbf{x}_{i-1}$ and $\mathbf{x}_{i}$ are neighbours in the lattice. Thus the distance between $\mathbf{x}$ and $\mathbf{y}$ in $\ZZ^2$ is equal to $\norm{\mathbf{x}-\mathbf{y}}$ \cite{Zerovnik1993}.
Each path $\mathbf{x}_0, \mathbf{x}_1,\dots,\mathbf{x}_k$ in $\ZZ^2$ gives rise to the oriented path $l(\mathbf{x}_0), l(\mathbf{x}_1), \ldots, l(\mathbf{x}_k)$ in $C_n(1, s)$. Note that even if the former is a shortest path in $\ZZ^2$, the latter is not necessarily a shortest path in $C_n(1, s)$.
Denote by $X$ the set of points of $\ZZ^2$ with label $0$. That is,
$$
X:=\left\{(x_1,x_2)\in\ZZ^2: l(x_1,x_2) = 0\right\}.
$$
Note that $X$ relies on $s$ implicitly. A \emph{basis} for $X$ is a set of two independent vectors $\left\{\mathbf{a},\mathbf{b}\right\}$ in $X$ such that any vector in $X$ is a linear combination of them with integer coefficients. The {\em parallelogram} \cite{Zerovnik1993} generated by a basis $\{\mathbf{a},\mathbf{b}\}$ is defined as
$$
[\mathbf{a},\mathbf{b}] := \left\{\mathbf{x} \in\ZZ^2: \mathbf{x}=\alpha \mathbf{a}+\beta\mathbf{b},0\leq\alpha,\beta \le 1\right\}.
$$
Note that its corner points, $\mathbf{0}, \mathbf{a},\mathbf{b}$ and $\mathbf{a+b}$, are in $X$.
Similarly, the {\em half-open parallelogram} generated by $\{\mathbf{a},\mathbf{b}\}$ is defined as
$$
[\mathbf{a},\mathbf{b}):=\left\{\mathbf{x} \in\ZZ^2: \mathbf{x}=\alpha \mathbf{a}+\beta\mathbf{b},0\leq\alpha,\beta<1\right\}.
$$
In what follows we use $d(i, j)$ to denote the distance in $C_n(1, s)$ between $i \in \ZZ_n$ and $j \in \ZZ_n$. It is observed in \cite{Zerovnik1993} that, for every $\mathbf{v}\in\ZZ^2$, we have
\begin{equation}
\label{eq:dl}
d\big(0,l(\mathbf{v})\big) := \norm{\mathbf{v}-X}:=\min\left\{\norm{\mathbf{v-x}}: \mathbf{x}\in X\right\}.
\end{equation}
\begin{lem} (\cite[Proposition 2]{Zerovnik1993})
\label{prop:conta}
Let $\{\mathbf{a},\mathbf{b}\}$ be a basis for $X$. Then $[\mathbf{a},\mathbf{b})$ has exactly $n$ points and each label in $\{0,1,\dots,n-1\}$ appears exactly once as $l(\mathbf{x})$ for some $\mathbf{x} \in [\mathbf{a},\mathbf{b})$.
\end{lem}
Thus the labelling $l$ induces a bijection between the points in $[\mathbf{a},\mathbf{b})$ and the nodes in $\ZZ_n$. This together with (\ref{eq:dl}) implies that for any $i \in \ZZ_n$,
$d(0, i) = \norm{\mathbf{v}_i-X}$, where $\mathbf{v}_i$ is the unique point in $[\mathbf{a},\mathbf{b})$ with $l(\mathbf{v}_i) = i$. The next lemma says that, if $\{\mathbf{a},\mathbf{b}\}$ is a packed basis, then $\norm{\mathbf{v}_i-X}$ is attained at a corner point of $[\mathbf{a},\mathbf{b}]$, where a basis $\{\mathbf{a,b}\}$ is \emph{packed}
\cite{Zerovnik1993} if it satisfies
\begin{equation}
\label{eq:packed}
\max\{\norm{\mathbf{a}},\norm{\mathbf{b}}\}\leq\min\{\norm{\mathbf{a-b}},\norm{\mathbf{a+b}}\}.
\end{equation}
\begin{lem}
(\cite[Lemma 2]{Zerovnik1993})
\label{lem:dist}
Let $\{\mathbf{a},\mathbf{b}\}$ be a packed basis for $X$. Then, for any $\mathbf{v} \in [\mathbf{a},\mathbf{b}]$, we have
$$
\norm{\mathbf{v} - X} = \min
\{{\norm{\mathbf{v} - \mathbf{x}}: \mathbf{x}} = \mathbf{0}, \mathbf{a}, \mathbf{b}, \mathbf{a} + \mathbf{b}\}.
$$
\end{lem}
\begin{lem}
\label{lem:pb}
The following hold:
\begin{itemize}
\item[\rm (a)] if $r\le q$ and $2r\le s + 1$, then $\{(s,-1),(r,q)\}$ is a packed basis for $X$;
\item[\rm (b)] if $r\le q$ and $2r\ge s + 1$, then $\{(s,-1),(r-s, q+1)\}$ is a packed basis for $X$;
\item[\rm (c)] if $r\ge q$ and $r + q \ge s + 1$, then $\{(s,-1),(r-s,q+1)\}$ is a packed basis for $X$.
\end{itemize}
\end{lem}
\begin{proof}
It can be shown (see \cite[pp.6]{Chen2005}) that any point in $X$ is of the form $i(s,-1)+j(r,q)$ for some integers $i$ and $j$. (In fact, for any $(x_1, x_2) \in X$, we have $x_1+x_2 s=kn$ for some integer $k$, and so $(x_1, x_2)=(kq-x_2)(s,-1)+k(r,q)$.) It follows that the pair $\{\mathbf{a,b}\}$ in each case is a basis for $X$. It remains to verify that $\{{\bf a}, {\bf b}\}$ satisfies (\ref{eq:packed}). Recall that $q \ge 2$ as $s < n/2$.
(a) Let ${\bf a} = (s, -1)$ and ${\bf b} = (r, q)$. Since $r\le q$, $2r\le s + 1$ and $q \ge 2$, we have $\norm{{\bf a}} = s+1, \norm{{\bf b}} = r+q, \norm{{\bf a} - {\bf b}} = s - r + q + 1$, $\norm{{\bf a}+{\bf b}} = s + r + q - 1$, and $\{{\bf a}, {\bf b}\}$ satisfies (\ref{eq:packed}).
(b) Let ${\bf a} = (s, -1)$ and ${\bf b} = (r - s, q + 1)$. Since $r\le q$ and $2r\ge s + 1$, we have $\norm{{\bf a}} = s+1, \norm{{\bf b}} = s - r+q+1, \norm{{\bf a} - {\bf b}} = 2s - r + q + 2$ and $\norm{{\bf a}+{\bf b}} = r + q$. Hence $\norm{{\bf a}} \le \norm{{\bf b}} \le \norm{{\bf a} + {\bf b}} \le \norm{{\bf a} - {\bf b}}$ and $\{{\bf a}, {\bf b}\}$ satisfies (\ref{eq:packed}).
(c) Let ${\bf a}$ and ${\bf b}$ be as in (b). Since $r\ge q$ and $r + q \ge s + 1$, the norms of ${\bf a}, {\bf b}, {\bf a} - {\bf b}, {\bf a}+{\bf b}$ are the same as in (b). Hence $\norm{{\bf b}} \le \norm{{\bf a}} \le \norm{{\bf a} + {\bf b}} \le \norm{{\bf a} - {\bf b}}$ and $\{{\bf a}, {\bf b}\}$ satisfies (\ref{eq:packed}).
\end{proof}
\begin{lem}
\label{lem:Ldist}
If $r\le q$ or $r + q\ge s + 1$, then
\begin{equation}
\label{eq:lb}
\sum_{i\in\ZZ_n}d(0,i)\geq \left\lfloor\frac{(s+1)^2}{2}\right\rfloor.
\end{equation}
\end{lem}
\begin{proof}
Since $r\le q$ or $r + q\ge s + 1$, one of the three cases in Lemma \ref{lem:pb} occurs, and in each case we have a packed basis $\{{\bf a}, {\bf b}\}$ for $X$ as given in Lemma \ref{lem:pb}. By Lemma \ref{prop:conta}, for any $i \in \ZZ_n$, there exists at least one $\mathbf{v}\in[\mathbf{a,b}]$ such that $l(\mathbf{v})=i$. Moreover, by Lemma \ref{lem:dist}, $d(0,i)=\min\{{\norm{\mathbf{v} - \mathbf{x}}: \mathbf{x}} = \mf{0}, \mf{a}, \mf{b}, \mf{a} + \mf{b}\}$. We now compute the sum of these distances $d(0,i)$ for $i$ in a certain subset of $\ZZ_n$.
\medskip\textsf{Case 1:}~$r\le q$ and $2r\le s+1$. In this case we have ${\bf a}=(s,-1),{\bf b}=(r,q)$ by Lemma \ref{lem:pb}. Set ${\bf v}_i=(i,0)$, ${\bf w}_i=({\bf a}+{\bf b})-(i,0)$, $\alpha_i=iq/n$, $\beta_i=i/n$, $\alpha_i'=1-iq/n$ and $\beta_i'=1-i/n$ for $1 \le i \le s$. Then $\alpha_i,\beta_i,\alpha_i',\beta_i'\in[0,1]$, ${\bf v}_i=\alpha_i{\bf a}+\beta_i{\bf b}$, ${\bf w}_i=\alpha_i'{\bf a}+\beta_i'{\bf b}$ and ${\bf v}_i, {\bf w}_i\in[{\bf a}, {\bf b}]$ for each $i$. Since $l({\bf v}_i)=i,l({\bf w}_i)=n-i$ and $s< n/2$, $l({\bf v}_1), \ldots, l({\bf v}_s), l({\bf w}_1), \ldots, l({\bf w}_s)$ are pairwise distinct. It can be verified that
$$
\norm{{\bf v}_i - {\bf 0}}=\norm{{\bf w}_i-({\bf a}+{\bf b})}=i,\quad \norm{{\bf v}_i-{\bf a}}=\norm{{\bf w}_i-{\bf b}}=s-i+1,
$$
$$
\norm{{\bf v}_i-({\bf a}+{\bf b})}=\norm{{\bf w}_i-{\bf 0}}=s+r-i+q-1,\quad \norm{{\bf v}_i-{\bf b}}=\norm{{\bf w}_i-{\bf a}}=|r-i|+q.
$$
Assume $1\leq i\leq \lfloor s/2\rfloor$. Since $r\leq q$, we have $|r-i| + q\ge i$ (as $i-r+q\ge i$ if $i>r$ and $(r-i)+q\ge r\ge i$ if $i\le r$) and $s+r-i+q-1 \ge s + 1 - i \ge i$. Therefore, $\norm{{\bf v}_i-X}=\norm{{\bf v}_i-{\bf 0}}$ and $\norm{{\bf w}_i-X}=\norm{{\bf w}_i-({\bf a}+{\bf b})}$. For $\lfloor s/2\rfloor< i\leq s$, it can be verified that
$\norm{{\bf v}_i-X}=\norm{{\bf v}_i-{\bf a}}$ and $\norm{{\bf w}_i-X}=\norm{{\bf w}_i-{\bf b}}$.
Therefore,
$$
\begin{array}{lll}
\sum_{i\in\ZZ_n}d(0,i) & \geq & \sum_{i=1}^{s}(\norm{{\bf v}_i-X}+\norm{{\bf w}_i-X}) \\[0.2cm]
& = & \sum_{i=1}^{\lfloor s/2 \rfloor} 2i + \sum_{\lfloor s/2 \rfloor + 1}^s 2(s-i+1) \\[0.2cm]
& \ge & \big \lfloor\frac{(s+1)^2}{2} \big \rfloor.
\end{array}
$$
\medskip
\textsf{Case 2:}~either $r\le q$ and $2r\ge s+1$, or $r\ge q$ and $r+q\ge s+1$. Then $r+q\ge s+1$, $2r\ge s+1$, and ${\bf a}=(s,-1), {\bf b}=(r-s,q+1)$ by Lemma \ref{lem:pb}.
Set ${\bf v}_i=(i,0)$, ${\bf w}_i=({\bf a}+{\bf b})-(i,0)$, $\alpha_i=i(q+1)/n$, $\beta_i=i/n$, $\alpha_i'=1-i(q+1)/n$, $\beta_i'=1-(i/n)$ for $1\le i\le s-1$ and ${\bf u}=(1,1)$. Then $\alpha_i,\beta_i,\alpha_i',\beta_i'\in[0,1]$, ${\bf v}_i=\alpha_i{\bf a}+\beta_i{\bf b}$, ${\bf w}_i=\alpha_i'{\bf a}+\beta_i'{\bf b}$, and ${\bf v}_i, {\bf w}_i \in[{\bf a}, {\bf b}]$ for each $i$. Moreover, ${\bf u}=((s-r+q+1)/n){\bf a}+((s+1)/n){\bf b} \in[{\bf a}, {\bf b}]$. Since $l({\bf v}_i)=i,\l({\bf w}_i)=n-i,l({\bf u})=s+1$ and $s<n/2$, $l({\bf v}_1),\dots, l({\bf v}_{s-1}), l({\bf w}_1), \dots, l({\bf w}_{s-1}) ,l({\bf u})$ are pairwise distinct.
It can be verified that
$$\norm{{\bf v}_i-{\bf 0}}=\norm{\mathbf{w}_i-({\bf a}+{\bf b})}=i,\quad \norm{{\bf v}_i-{\bf a}}=\norm{{\bf w}_i-{\bf b}}=s-i+1,$$
$$\norm{{\bf v}_i-({\bf a}+{\bf b})}=\norm{{\bf w}_i-{\bf 0}}=|r-i|+q, \quad \norm{{\bf v}_i-{\bf b}}=\norm{{\bf w}_i-{\bf a}}=s-r+i+q+1.$$
Assume $1\leq i\leq \lfloor s/2\rfloor$. Since $r+q\ge s+1$, we have $ r-i+q\ge s-i+1\ge i$ and $ s-r+i+q+1\ge i$. Hence $\norm{\mathbf{v}_i-X}=\norm{\mathbf{v}_i-{\bf 0}}$ and $\norm{{\bf w}_i-X}=\norm{{\bf w}_i-({\bf a}+{\bf b})}$. If $\lfloor s/2\rfloor< i\leq s-1$, then $\norm{\mathbf{v}_i-X}=\norm{\mathbf{v}_i-{\bf a}}$ and $\norm{\mathbf{w}_i-X}=\norm{\mathbf{w}_i-{\bf b}}$. Note that $\norm{{\bf u}-X}=2.$
Therefore,
$$
\begin{array}{lll}
\sum_{i\in\ZZ_n}d(0,i) & \geq & \sum_{i=1}^{s-1}(\norm{\mathbf{v}_i-X}+\norm{\mathbf{w}_i-X})+\norm{{\bf u}-X} \\ [0.2cm]
& = & \sum_{i=1}^{\lfloor s/2 \rfloor} 2i + \sum_{\lfloor s/2 \rfloor + 1}^{s-1} 2(s-i+1) + 2 \\[0.2cm]
& \ge & \big\lfloor \frac{(s+1)^2}{2}\big\rfloor.
\end{array}
$$
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:lwr3}]
Suppose $r \le q$ or $r + q \ge s + 1$. Since $C_{n}(1, s)$ is vertex-transitive, we have $\sum_{(i, j) \in \ZZ_n \times \ZZ_n} d(i, j) = n \sum_{i \in \ZZ_n} d(0, j)$. This together with (\ref{eq:trvl}) and (\ref{eq:lb}) implies (\ref{eq:lwr3}).
\end{proof}
As shown in the proof of Lemma \ref{lem:Ldist}, we obtained (\ref{eq:lb}) by computing the total distance from the vertex $0$ to $2s$ or $2s-1$ other vertices. This implies that the difference between the two sides of (\ref{eq:lb}) is small if and only if $s$ is close to $n/2$.
The sum of the distances can be precisely computed when $s= \sqrt{n}$, leading to the following better lower bound than Theorem \ref{thm:lwr3} in this special case.
\begin{lem}
\label{lem:ssqrtn}
If $s= q=\sqrt{n}$, then
\[
\pi(C_n(1,s)) \ge \frac{\sqrt{n}(n-1)}{4}.
\]
\end{lem}
\begin{proof}
By Lemma \ref{lem:pb}, $\{ (\sqrt{n},-1), (0,\sqrt{n})\}$ is a packed basis for $X$ in this case. One can see that the following is a closest corner point to $(i,j)$:
\begin{itemize}
\item[\rm (a)] $(0,0)$, if $0\le i \le \lfloor \sqrt{n}/2 \rfloor$ and $0\le j\le \lfloor (\sqrt{n}-1)/2 \rfloor$;
\item[\rm (b)] $(\sqrt{n},-1)$, if $\lfloor \sqrt{n}/2 \rfloor +1 \le i\le \sqrt{n}-1 $ and $0\le j \le \lfloor \sqrt{n} /2 \rfloor-1$;
\item[\rm (c)] $(0,\sqrt{n})$, if $0 \le i \le \lfloor (\sqrt{n}-1)/2 \rfloor$ and $\lfloor (\sqrt{n}+ 1)/2 \rfloor \le j \le \sqrt{n} -1$ and
\item[\rm (d)] $( \sqrt{n},\sqrt{n}-1)$, if $\lfloor (\sqrt{n}+1)/2 \rfloor \le i \le \sqrt{n}-1$ and $\lfloor \sqrt{n}/2 \rfloor \le j \le \sqrt{n} -1$.
\end{itemize}
Since every point $(i, j)$ in $X$ appears in exactly one of these four cases, we have
\[
\sum_{k\in\ZZ_n}d(0,k) = \sum_{i=0}^{\lfloor \sqrt{n}/2 \rfloor}\sum_{j=0}^{\lfloor (\sqrt{n}-1)/2 \rfloor} (i+j) + \sum_{i=\lfloor \sqrt{n}/2 \rfloor+1}^{\sqrt{n}-1 } \sum_{j=0}^{\lfloor \sqrt{n}/2 \rfloor-1} (\sqrt{n}-i+j+1) +
\]
\[
\sum_{i=0}^{\lfloor (\sqrt{n}-1)/2 \rfloor}\sum_{j=\lfloor (\sqrt{n}+1)/2 \rfloor}^{\sqrt{n}-1 } (i+\sqrt{n}-j) + \sum_{i=\lfloor (\sqrt{n}+1)/2 \rfloor}^{\sqrt{n}-1 } \sum_{j=\lfloor \sqrt{n} /2 \rfloor}^{\sqrt{n}-1 } (\sqrt{n}-i+\sqrt{n}-1-j)= \frac{\sqrt{n}(n-1)}{2}.
\]
The result then follows from (\ref{eq:trvl}) and vertex-transitivity of $ C_n(1,s)$.
\end{proof}
\section{A routing scheme, and proof of Theorem \ref{thm:tr1}}
\label{sec:ub}
In this section we give a specific routing scheme for $C_n(1,s)$ which yields the required upper bounds on the forwarding indices of $C_n(1,s)$. The same routing will be used in the next section to give upper bounds on the optical indices of $C_n(1,s)$. We will use the words `link' and `arc' interchangeably and we call a link of $C_n(1,s)$ of the form $(x,x\oplus1)$ ($(x,x\ominus 1)$, respectively) a {\em clockwise} ({\em anticlockwise}, respectively) {\em ring link}, and a link of the form $(x,x\oplus s)$ ($(x,x\ominus s)$, respectively) a {\em clockwise} ({\em anticlockwise}, respectively) {\em skip link}. We define a routing as follows.
\begin{const}
\label{const}
{\em
Define
\begin{equation}
\label{eq:R}
{\cal R} := \{P_{x, y}: x, y \in \ZZ_n, x \ne y\},
\end{equation}
where $P_{x, y}$ is the path in $C_n(1,s)$ from $x$ to $y$ specified as follows.
\begin{enumerate}
\item \label{step:s1}
For $d = 1, \ldots, \lfloor n/2\rfloor$, say, $d=is+j$ for some $i,j$ with $0\le i\le \lfloor q/2\rfloor$ and $0\le j\le s-1$,
\begin{enumerate}[(a)]
\item if $j \leq \lfloor s/2\rfloor $, then define $P_{0,d}: 0,s,2s,\dots,is,is+1, is + 2, \dots,is+j$;
\item if $j> \lfloor s/2\rfloor $, then define $P_{0,d}: 0,s,2s,\dots,(i+1)s,(i+1)s-1,(i+1)s-2,\dots,(i+1)s-(s-j)$.
\end{enumerate}
\item \label{step:s3}
For $d = \lfloor n/2\rfloor +1, \ldots, n-1$, letting $P_{0, n-d}: v_1,v_2,\dots,v_k$ be the path from $0$ to $n-d$ constructed in Step 1, define $P_{0,d}: v_1, n-v_2,\dots, n-v_k$.
\item \label{step:s4}
For $1\leq x, y\leq n-1$ with $x \ne y$, letting $v_1,v_2,\dots,v_k$ denote the path $P_{0,y\ominus x}$ from $0$ to $y\ominus x$ constructed in Step 1 or 2,
define $P_{x,y}: x\oplus v_1,x\oplus v_2,\dots, x\oplus v_k$.
\end{enumerate}
}
\end{const}
The routing $\cal R$ constructed above is symmetric in the sense that, for any $x, y, k \in \ZZ_n$ with $x \ne y$, $P_{x\oplus k, y\oplus k}$ is the path obtained by adding $k$ to each node of $P_{x, y}$. We say that $P_{x\oplus k, y\oplus k}$ is obtained from {\em translation} of $P_{x, y}$ by $k$. This feature is crucial in the following computation of ${\overrightarrow{\pi}}(C_n(1,s),{\cal R})$. Denote
$$
\Delta := \left\{
\begin{array}{ll}
\frac{s}{4} + \frac{1}{2} \left\lfloor\frac{r}{2}\right\rfloor \left(s - \left\lfloor\frac{r+2}{2}\right\rfloor\right),\;\; \mbox{if $s$ is even}\\ [0.3cm]
\frac{1}{2} \left\lfloor\frac{r+1}{2}\right\rfloor \left(s - \left\lfloor\frac{r+1}{2}\right\rfloor\right),\;\; \mbox{if $s$ is odd}.
\end{array}
\right.
$$
One can verify that the four numbers involved in the next lemma are integers.
\begin{lem}
\label{lem:ub}
\begin{itemize}
\item[\rm (a)] If $q$ is even, then
$$
{\overrightarrow{\pi}}(C_n(1,s),{\cal R}) =
\max\left\{ \frac{q}{4} \BL{s^2}{2} + \frac{1}{2}\left\lfloor \frac{r}{2}\right\rfloor \BL{r+2}{2},
\frac{q^2s}{8} + \frac{q}{2} \left(\left\lfloor \frac{r}{2}\right\rfloor + \frac{\epsilon(s) }{2}\right) \right\}.
$$
\item[\rm (b)] If $q$ is odd, then
$$
{\overrightarrow{\pi}}(C_n(1,s),{\cal R}) =
\max\left\{ \frac{q}{4}\BL{s^2}{2} +\Delta,
\frac{(q^2-1)s}{8} + \frac{q+1}{2}\left\lfloor\frac{r+\epsilon(s)}{2}\right\rfloor
\right\}.
$$
\end{itemize}
\end{lem}
\begin{proof}
Observe that if a path $P_{x, y}$ in $\cal R$ passes through a clockwise ring link $(v, v\oplus1)$, then the path $P_{x\oplus (v'- v), y\oplus (v' - v)}$ (which is also in $\cal R$) passes through the clockwise ring link $(v', v'\oplus 1)$. Hence the loads on all clockwise ring links under $\cal R$ are equal. Similarly, all anticlockwise ring links have the same load, all clockwise skip links have the same load, and all anticlockwise skip links have the same load. So the load on each clockwise ring link (skip link, respectively) is the total number of clockwise ring links (skip links, respectively) used by paths of $\cal R$ divided by $n$. The same can be said for anticlockwise ring or skip links.
On the other hand, for any fixed $x\in\ZZ_n$, the paths $P_{x,y}$, $x \ne y=0,\dots,n-1$, use the same number of ring (skip, respectively) links as the paths $P_{0,d}$, $d=1,\dots,n-1$, because $P_{x,y}$ is the translation of $P_{0,y\ominus x}$ by $x$ and therefore there is a bijection between these two sets of paths. Since there are $n$ translations for any path $P_{0,d}$, $d=1,\dots,n-1$, we conclude that the load on any ring (skip, respectively) link in each direction is the total number of used ring (skip, respectively) links in that direction by paths $P_{0,d}$, $d=1,\dots,n-1$.
\medskip
\textsf{Claim 1:}~The maximum load on ring links under $\cal R$ is equal to the number of clockwise ring links used when $q$ is even, and anticlockwise ring links used when $q$ is odd, by paths $P_{0,d}$, $d=1,\dots,n-1$.
\medskip
\textsf{Claim 2:}~The maximum load under $\cal R$ on skip links is equal to the number of clockwise skip links used by paths $P_{0,d}$, $d=1,\dots,n-1$.
\medskip
\textit{Proof of Claims 1 and 2.}~For any path $P_{0,d}$ in ${\cal R}$, where $d < n/2 $, the path $P_{0,n - d}$ is in ${\cal R}$ and is distinct from $P_{0,d}$. Moreover, if $(u, v)$ is a link in one of these two paths, then $(n\ominus u,n\ominus v)$ is a link in the other (with opposite direction). If $n$ is odd, then $d\ne n - d$ for all $d\in\ZZ_n$, and so the number of clockwise ring links (skip links, respectively) used is equal to the number of anticlockwise ring links (skip links, respectively) used by the paths $P_{0,d}$.
Assume $n$ is even. If $d=n/2$, then $P_{0,d}$ and $P_{0,n - d}$ are identical, and this path does not use any anticlockwise skip link. Hence $P_{0,n/2}$ uses fewer anticlockwise skip links than clockwise skip links. So the maximum load on skip links is equal to the number of clockwise skip links used by the paths $P_{0,d}$.
If $q$ is even, then $n/2=q s /2 + r /2$; in this case $P_{0,n/2}$ uses $r/2$ clockwise ring links, and so it uses fewer anticlockwise ring links than clockwise ring links. Thus the maximum load on ring links is equal to the number of clockwise ring links used by the paths $P_{0,d}$.
If $q$ is odd, then $n/2 = (q-1)s /2 + (s+r)/2$; in this case $P_{0,n/2}$ uses $ (s-r)/2$ anticlockwise ring links, and so it uses fewer clockwise ring links. Hence the maximum load on ring links is equal to the number of anticlockwise ring link used by the paths $P_{0,d}$. This completes the proof of Claims 1 and 2.
\qed
We now count the number of links in $C_n(1,s)$ used by paths $P_{0,d}$, $d=1,\dots,n-1$, for even $q$ and odd $q$ separately.
\medskip
\textsf{Ring links:}~We first count the number of clockwise (anticlockwise, respectively) ring links used by paths $P_{0,d}$ when $q$ is even (odd, respectively), where $P_{0,d}$ is defined in Construction \ref{const} in Step 1 when $d=is+j \le\lfloor n/2\rfloor $ and in Step 2 when $d=n-(is+j) >\lfloor n/2\rfloor$, where $0\le i\le \lfloor q/2\rfloor$ and $0\le j\le s-1$.
\medskip
\textsf{Case 1:}~$q$ is even. If $d\le \lfloor n/2\rfloor$, then $is+j \le qs/2 +\lfloor r/2\rfloor$; if $d >\lfloor n/2\rfloor$, then $d=n-(is+j)$ and so $is+j<qs/2+\lceil r/2 \rceil$.
When $d\le \lfloor n/2\rfloor$, by Step \ref{step:s1}, the path $P_{0,d}$ uses some clockwise ring link if and only if either $0 \le i \le q/2-1$ and $0 \le j \le \lfloor s/2\rfloor$, or $i=q/2$ and $0 \le j \le \lfloor r/2\rfloor$. When $d>\lfloor n/2\rfloor $, by Steps \ref{step:s1}(b) and \ref{step:s3}, $P_{0,d}$ uses some clockwise ring link if and only if $1 \le i \le q/2$ and $\lfloor s/2\rfloor+1 \le j \le s-1$. Moreover, $P_{0,d}$ uses $j$ ($s-j$, respectively) clockwise ring links if $d\le \lfloor n/2\rfloor$ ($d> \lfloor n/2\rfloor$, respectively). Therefore, the total number of clockwise ring links used by the paths $P_{0,d}$, $d=1,\dots,n-1$, is equal to
\begin{equation}
\label{eq:even1}
\sum_{i=0}^{q/2-1}\sum_{j=0}^{\lfloor s/2\rfloor} j + \sum_{i=q/2}^{q/2}\sum_{j=0}^{\lfloor r/2\rfloor} j + \sum_{i=1}^{q/2}\sum_{j=\lfloor s/2\rfloor+1}^{s-1} (s-j) =
\frac{q}{4} \left\lfloor \frac{s^2}{2} \right\rfloor + \frac{1}{2}\left\lfloor \frac{r}{2}\right\rfloor\left\lfloor\frac{r+2}{2}\right\rfloor.
\end{equation}
\medskip
\textsf{Case 2:}~$q$ is odd. If $d\le \lfloor n/2\rfloor$, then $is +j \le (q-1)s/2 + \lfloor (s+r)/2\rfloor$; if $d >\lfloor n/2\rfloor$, then $d=n-(is+j)$ and so $is+j < (q-1)s/2 + \lceil (s+r)/2 \rceil$.
When $d\le \lfloor n/2\rfloor$, by Step \ref{step:s1}(b), $P_{0,d}$ uses some anticlockwise ring link if and only if either $1 \le i \le (q-1)/2$ and $\lfloor s/2\rfloor+1 \le j \le s-1$, or $i= (q+1)/2$ and $\lfloor s/2\rfloor+1 \le j \le \lfloor (s+r)/2\rfloor$. When $d > \lfloor n/2\rfloor$, by Steps \ref{step:s1}(a) and \ref{step:s3}, $P_{0,d}$ uses some anticlockwise ring link if and only if $0 \le i \le (q-1)/2$ and $1 \le j \le \lfloor s/2\rfloor$.
The path $P_{0,d}$ uses $s-j$ ($j$, respectively) anticlockwise ring links if $d\le \lfloor n/2\rfloor$ ($d> \lfloor n/2\rfloor$, respectively).
Therefore, the total number of anticlockwise ring links used by the paths $P_{0,d}$, $d=1,\dots,n-1$, is given by
\begin{equation}
\label{eq:odd1}
\sum_{i=1}^{(q-1)/2}\sum_{j=\lfloor s/2\rfloor+1}^{s-1} (s-j) + \sum_{i=(q+1)/2}^{(q+1)/2}\sum_{j=\lfloor s/2\rfloor+1}^{\lfloor (s+r)/2\rfloor} (s-j) + \sum_{i=0}^{(q-1)/2}\sum_{j=1}^{\lfloor s/2\rfloor}j =
\frac{q}{4} \left\lfloor \frac{s^2}{2} \right\rfloor + \Delta.
\end{equation}
\medskip
\textsf{Skip links:}~
Now we evaluate the load on clockwise skip links. Note that $P_{0,d}$ uses clockwise skip links if $d\le \lfloor n/2\rfloor $.
\medskip
\textsf{Case 3:}~$q$ is even. In this case $P_{0,d}$ uses exactly $k$ clockwise skip links if and only if either (i) $i=k$, $0\le j \le \lfloor s/2\rfloor $, if $1\le k \le q/2-1$; (ii) $i=k$, $0 \le j \le \lfloor r/2\rfloor$, if $k=q/2$; or (iii) $i=k-1$, $\lfloor s/2\rfloor +1 \le j \le s-1$, if $1\le k \le q/2$. So the total number of clockwise skip links used by the paths $P_{0,d}$, $d=1,\dots,n-1$, is equal to
\begin{equation}
\label{eq:even2}
\sum_{k=1}^{q/2-1}k \left(\left\lfloor \frac{s}{2}\right\rfloor +1\right) + \sum_{k=q/2}^{q/2} k\left(\left\lfloor \frac{r}{2}\right\rfloor+1\right) + \sum_{k=1}^{q/2} k\left(s - \left\lfloor \frac{s}{2}\right\rfloor -1\right) = \\[.2cm]
\frac{q^2s}{8} + \frac{q}{2} \left(\left\lfloor \frac{r}{2}\right\rfloor + \frac{\epsilon(s) }{2}\right) .
\end{equation}
\medskip
\textsf{Case 4:}~$q$ is odd. By Step \ref{step:s1}, $P_{0,d}$ uses exactly $k$ clockwise skip links if and only if either (i) $i=k$, $0 \le j \le \lfloor s/2\rfloor$, if $1\le k \le (q-1)/2$; (ii) $i=k-1$, $\lfloor s/2\rfloor +1\le j \le s-1$, if $1\le k \le (q-1)/2$; or (iii) $i=k-1, \lfloor s/2\rfloor+1 \le j \le \lfloor (s+r)/2\rfloor$, if $k=(q+1)/2$. Thus the total number of clockwise skip links used by the paths $P_{0,d}$, $d=1,\dots,n-1$, is equal to
\begin{equation}
\label{eq:odd2}
\sum_{k=1}^{(q-1)/2} k \left\lfloor \frac{s+2}{2}\right\rfloor + \sum_{k=1}^{(q-1)/2} k \left\lceil \frac{s-2}{2} \right\rceil + \sum_{k=(q+1)/2}^{(q+1)/2} k \left\lfloor\frac{r+\epsilon(s)}{2}\right\rfloor
=\frac{(q^2-1)s}{8}+ \frac{q+1}{2}\left\lfloor\frac{r+\epsilon(s)}{2}\right\rfloor.
\end{equation}
Using (\ref{eq:even1})-(\ref{eq:odd2}), Claims 1 and 2 imply the required results immediately.
\end{proof}
By comparing the two terms in each case of Lemma \ref{lem:ub}, we can identify the maximum term for different ranges of $s$, which is presented in the following lemma.
\begin{lem}
\label{lem:ub1}
The following hold:
\begin{itemize}
\item[\rm (a)] if $q$ is even and $2 \le s \le \sqrt{n-1}$, then
\begin{equation}
\label{eq:ub}
{\overrightarrow{\pi}}(C_n(1,s),{\cal R}) \le \frac{q(n+r+2\epsilon(s))}{8};
\end{equation}
\item[\rm (b)] if $q$ is odd and $2 \le s \le \sqrt{n} -1$, then
\begin{equation}
\label{eq:ub1}
{\overrightarrow{\pi}}(C_n(1,s),{\cal R}) \le \frac{q( n+ r+2\epsilon(s)) +s }{8};
\end{equation}
\item[\rm (c)] if $q=s=\sqrt{n}$, then the loads on all links of $C_n(1,s)$ are equal and
\begin{equation}
\label{eq:ub2}
{\overrightarrow{\pi}}(C_n(1,s),{\cal R}) = \frac{\sqrt{n}\ (n - \epsilon(s))}{8};
\end{equation}
\item[\rm (d)] if $q$ is even and $s \ge \sqrt{n}+1$, then
\begin{equation}
\label{eq:ub3}
{\overrightarrow{\pi}}(C_n(1,s),{\cal R}) \le \frac{sn+r-\epsilon(s)q}{8};
\end{equation}
\item[\rm (e)] if $q$ is odd and $s \ge \sqrt{n}$, then
\begin{equation}
\label{eq:ub4}
{\overrightarrow{\pi}}(C_n(1,s),{\cal R}) \le \frac{s (n + r+ 2) - \epsilon(s)q}{8}.
\end{equation}
\end{itemize}
\end{lem}
Moreover, when $\sqrt{n}$ is not an integer, if $s = \lfloor\sqrt{n}\rfloor$ and $q$ is odd or $s = \lfloor\sqrt{n}+1\rfloor$ and $q$ is even, then the greater term in each case in Lemma \ref{lem:ub} relies on $r$. However, the two terms are almost equal for sufficiently large $n$. When $\sqrt{n}$ is an odd integer and $q=s=\sqrt{n}$, the conditions in cases (c) and (e) are satisfied simultaneously. Although the right hand sides of (\ref{eq:ub2}) and (\ref{eq:ub4}) are equal in this special case, the result in case (c) is slightly stronger as we have equality in (\ref{eq:ub2}).
\begin{proof}[Proof of Theorem \ref{thm:tr1}]
As mentioned in Section \ref{subsec:two}, the lower bounds in (\ref{eq:tf1})-(\ref{eq:tf3}) follow from (\ref{eq:pi}) and Lemmas \ref{lem:lwr1}, \ref{lem:lwr2} and \ref{lem:ssqrtn} immediately. The upper bounds in (\ref{eq:tf1})-(\ref{eq:tf3}) follow from Lemma \ref{lem:ub1} and the fact that ${\overrightarrow{\pi}}(C_n(1,s)) \le {\overrightarrow{\pi}}(C_n(1,s),{\cal R})$. In fact, if $2 \le s \le \sqrt{n} -1$, then (\ref{eq:ub}) or (\ref{eq:ub1}) applies. By (\ref{eq:ub}) and (\ref{eq:ub1}), we obtain ${\overrightarrow{\pi}}(C_n(1,s)) \le (q( n+ r+2\epsilon(s))/8)+(s/8) = ((n-r)(n+r+2)+s^2)/8s$. If $s=\sqrt{n}$, then $q=\lfloor n/s \rfloor =\sqrt{n}$ and by (\ref{eq:ub2}) we obtain ${\overrightarrow{\pi}}(C_n(1,s)) \le (\sqrt{n} (n - \epsilon(s) ) / 8$.
If $\sqrt{n}+1 \le s < n/2$, then by (\ref{eq:ub3}) and (\ref{eq:ub4}) we have ${\overrightarrow{\pi}}(C_n(1,s)) \le (s(n +r +2)-\epsilon(s)q)/8 = (s^2(n+r+2)-\epsilon(s)(n-r))/8s$.
\end{proof}
\section{Proof of Theorem \ref{thm:tr2}}
\label{sec:wave}
The lower bounds in (\ref{eq:thw1})-(\ref{eq:thw3}) follow from (\ref{eq:piw}) and Theorem \ref{thm:tr1} immediately. Let $\cal R$ be the routing defined in (\ref{eq:R}). Since ${\overrightarrow{w}}(C_n(1,s)) \le {\overrightarrow{w}}(C_n(1,s),{\cal R})$ and $w(C_n(1,s)) \le w(C_n(1,s),{\cal R})$, it suffices to prove the upper bounds for ${\overrightarrow{w}}(C_n(1,s),{\cal R})$ and $w(C_n(1,s),{\cal R})/2$. In the following we give an arc-conflict-free colouring of ${\cal R}$ and compute the number of colours used. This number gives the required upper bound for ${\overrightarrow{w}}(C_n(1,s), {\cal R})$.
By Construction \ref{const}, the unique path in $\cal R$ connecting $x$ to $y=x\oplus (is+j)$ is $P_{x,y} $, which we denote by $P_{ij}^x$ in the rest of this proof, where $|i| \le \lceil q/2\rceil$, $|j| \le \lfloor s/2\rfloor$ and $|is+j|\le n/2$. In other words, $P_{ij}^x$ connects $x$ to $y$ by $i$ successive skip links followed by $j$ successive ring links, where negative $i$ and $j$ stand for anticlockwise skip and ring links, respectively.
Given $i$ and $j$, set
$$
\alpha := |j|/\gcd(s,|j|),\;\, \beta := s/\gcd(s,|j|),
$$
where $\gcd(s,|j|)$ is the greatest common divisor of $s$ and $|j|$. Then $\alpha$ is the smallest positive integer such that $|j|$ divides $\alpha s$. Note that $\alpha s=\beta|j|$.
\begin{const}
\label{colouring}
{\em
We define a colouring $f: {\cal R} \rightarrow \ZZ^3$ by using elements of $\ZZ^3$ as colours.
\begin{enumerate}
\item
If $j>0$ and $|i| \le j$, let $x_{\alpha}=\lfloor x/\alpha s\rfloor$.
\begin{enumerate}[(a)]
\item
If $ x \le n/2$, define $f(P^x_{ij})=(\epsilon(x_{\alpha}) j +(x+\lfloor x_{\alpha}/2\rfloor \mod{j}),i,j)$;
\item
if $x > n/2$, define $f(P^x_{ij})=( (2 + \epsilon(x_{\alpha})) j +(x+\lfloor x_{\alpha}/2\rfloor \mod{j}),i,j)$.
\end{enumerate}
\item
If $|i| > j\geq0$, let $x_0=x \mod s$.
\begin{enumerate}[(a)]
\item
If $x < \lfloor q/2\rfloor s$ and $x_0 \leq s - j$, define $f(P^x_{ij})=(x_0+\lfloor x/s\rfloor \mod{|i|},i,j)$;
\item if $x <(\lfloor q/2\rfloor-1)s$ and $x_0 > s - j$, define $f(P^x_{ij})=( i + (x_0 +\lfloor x/s\rfloor \mod{|i|}), i , j)$;
\item
if $(\lfloor q/2\rfloor - 1) s \le x < \lfloor q/2\rfloor s$ and $x_0 > s - j$, define $f(P^x_{ij})=( 2i + x_0 + j - s , i , j)$;
\item
if $\lfloor q/2\rfloor s \le x \le n - j $ and $x_0 \leq s-j$, define $f(P^x_{ij})=(i+(x_0+\lfloor x / s\rfloor+ s-q-r -1 \mod{|i|}),i,j)$;
\item
if $\lfloor q/2\rfloor s \le x \le n - j $ and $x_0 > s-j$, define $f(P^x_{ij})=( x_0 +\lfloor x / s \rfloor -q-r \mod{|i|}, i , j)$;
\item
if $ x > n - j $, define $f(P^x_{ij})=( 2i + x + j - n, i , j)$.
\end{enumerate}
\item If $j<0$, define $f(P^x_{ij})=f(P^x_{(-i)(-j)})$.
\end{enumerate}
}
\end{const}
By the definition above, $f(P^x_{i j}) \ne f(P^{y}_{k l})$ if $k \ne i$ or $l \ne j$ when $jl \ge 0$, and if $k \ne -i$ or $l \ne -j$ when $jl < 0$.
Since $P^x_{ij}$ and $P^y_{(-i)(-j)}$ have no common link in the same direction, to prove that $f$ is arc-conflict-free, it suffices to verify that $P^x_{ij}$ and $P^y_{ij}$ do not have any common link if $f(P^x_{i j}) = f(P^{y}_{i j})$, or equivalently $f(P^x_{i j}) \ne f(P^{y}_{i j})$ if $P^x_{ij}$ and $P^y_{ij}$ have a common link.
Fix $i$ and $j$. Assume $x < y$. Note that $P^x_{i j}$ and $P^y_{i j}$ share a skip link if and only if $y \ominus x = h s$ for some $h$ with $0< |h| < |i|$, and they share a ring link if and only if $y \ominus x < |j|$.
\medskip
\textsf{Case 1:}~$j>0$ and $|i|\leq j$. Assume $f(P^x_{i j})=f(P^y_{i j})$. Then by Step 1, $x_\alpha$ and $y_\alpha$ have the same parity and either $x,y \le n/2$ or $x,y> n/2$. So if $P^x_{ij}$ and $P^y_{ij}$ share a ring link, then $y - x<j$, and if they share a skip link, then $y-x= h s$ for some $h$ with $0<h<|i|$. In the following we show that neither of these can happen, and therefore $P^x_{ij}$ and $P^y_{ij}$ cannot have any common link.
\medskip
\textsf{Subcase 1.1:}~$x_\alpha=y_\alpha$. We have $x+\lfloor x_\alpha/2\rfloor \equiv y +\lfloor y_\alpha/2\rfloor\mod{j}$ as $f(P^x_{ij})=f(P^y_{ij})$. So $y=x+\gamma j$ as $x_\alpha=y_\alpha$, where $\gamma\ge 1$.
Hence $P^x_{ij}$ and $P^y_{ij}$ do not share any ring link as $y - x\ge j$. If they share a skip link, then $y = x+hs$, $0<h< |i|$, which together with $y = x+\gamma j$ gives $h s=\gamma j$. Since $x_\alpha=y_\alpha$, we have $y - x <\alpha s$ and so $h<\alpha$. The latter inequality together with $hs=\gamma j$ contradicts the choice of $\alpha$. Hence $P^x_{ij}$ and $P^y_{ij}$ do not share any skip link.
\medskip
\textsf{Subcase 1.2:}~$x_\alpha\ne y_\alpha$. We have $x=x_\alpha \alpha s+l_x s+x_0$ and $y=y_\alpha \alpha s+ l_y s+y_0$, where $0\le l_x,l_y\le \alpha-1$, $x_0= x\mod s$ and $y_0= y \mod s$.
Since $x_\alpha$ and $y_{\alpha}$ have the same parity, we have $y_\alpha-x_\alpha\ge 2$ and so $y - x = (y_\alpha - x_\alpha) \alpha s + (l_y - l_x)s + y_0 - x_0 > \alpha s =\beta j \ge j$ as $|l_y - l_x|\le \alpha - 1$ and $|y_0 - x_0| < s$.
So $P^x_{ij}$ and $P^y_{ij}$ do not share any ring link.
Suppose by way of contradiction that $P^x_{ij}$ and $P^y_{ij}$ have a common skip link. Then $y = x + h s$ with $h=\alpha + t$, for a positive integer $t$, as $y_\alpha - x_\alpha \ge 2$. Also $y_\alpha\alpha s+ l_y s =x_\alpha\alpha s+ l_x s +h s$ as $y_0=x_0$. Since $f(P^x_{ij}) = f(P^y_{ij})$, we have $ x + \lfloor x_{\alpha}/2\rfloor \equiv y +\lfloor y_\alpha/2\rfloor \equiv x +hs+\lfloor y_\alpha/2\rfloor\mod j$. Hence $(y_\alpha - x_\alpha)/2 \equiv -h s \equiv -(\alpha s+t s) \equiv - t c \mod j$ as $\alpha s = \beta j$, where $ c = s \mod{j}$. If $c=0$, then $y_\alpha-x_\alpha\ge 2j$ as $y_\alpha>x_\alpha$ and $(y_\alpha-x_\alpha)\alpha s \ge 2j\alpha s$.
If $c\ne0$, then $(y_\alpha-x_\alpha)/2=k j - t c=\gcd(c,j)(k a_1 - t a_2)$, where $\gcd(a_1,a_2)=1$ and $k$ is an integer. So $y_\alpha - x_\alpha \ge 2 \gcd(c,j)$ as $y_\alpha > x_\alpha$. Since $\alpha = j / \gcd(s,j)$ and $\gcd(s \mod j,j)=\gcd(s,j)$, we have $\alpha = j / \gcd(c,j)$, $c \ne 0$. So $ (y_\alpha - x_\alpha) \alpha s \ge 2 \gcd(c,j) \alpha s = 2 j s$. Since $(y_\alpha - x_\alpha) \alpha s = hs + (l_x - l_y) s$, we have $hs + (l_x - l_y) s \ge 2 j s$, and so $ h s \ge 2 j s + (l_y - l_x) s \ge 2 j s + (1-\alpha )s=s + (2s - \beta) j \ge (j+1) s > |i| s$ as $\beta\le s$ and $j\ge |i|$, contradicting the fact $h < |i|$. Hence $P^x_{ij}$ and $P^y_{ij}$ have no common skip link.
\medskip
\textsf{Case 2:}~$|i|>j\geq0$. We show that $P^x_{ij}$ and $P^y_{ij}$ are assigned distinct colours if they share a link. We first assume that they share a ring link, so that either $y - x \le j - 1$ or $x - y + n \le j - 1$. Denote by $f(P^x_{ij})_1$ the first coordinate of $f(P^x_{ij})$.
\medskip
\textsf{Subcase 2.1:}~$y - x \le j - 1$. Since $j\le \lfloor s/2\rfloor$, either $\lfloor y /s \rfloor=\lfloor x/s \rfloor+1$ or $\lfloor y /s \rfloor=\lfloor x /s \rfloor$.
If $\lfloor y /s \rfloor=\lfloor x /s \rfloor + 1$, then $y_0 \le j$ and $x_0 > s - j$ as $y - x \le j-1$, which implies that $y$ and $x$ respectively satisfy either (a) and (b), or (d) and (c), or (d) and (e), or (f) and (e) in Step 2, and so $f(P^x_{ij})_1$ and $f(P^y_{ij})_1$ differ by at least $i$.
Now assume $\lfloor y/s \rfloor=\lfloor x/s \rfloor$. If $x_0 \le s-j$ and $y_0 > s-j$, then $f(P^x_{ij})_1$ and $f(P^y_{ij})_1$ differ by at least $i$; otherwise
$x_0 + \lfloor x/s \rfloor \not\equiv y_0 + \lfloor y/s \rfloor \mod i$ as $y_0 - x_0 \le j-1 < |i|$. So $f(P^x_{ij})\ne f(P^y_{ij})$.
\medskip
\textsf{Subcase 2.2:}~$x - y + n \le j - 1$. In this case we have $0 \le x < j$ and $n - j < y < n$, and so $f(P^x_{ij})_1$ and $f(P^y_{ij})_1$ differ by at least $2i$ by (a) and (f) in Step 2.
Now we assume that $P^x_{ij}$ and $P^y_{ij}$ share a skip link, so that either $y=x+hs$ or $x= y +hs-n$, where $0< h <|i|$.
\medskip
\textsf{Subcase 2.3:}~$y = x+hs$. So $y_0=x_0$ and $\lfloor y/s \rfloor - \lfloor x/s \rfloor = h$. If both $x$ and $y$ satisfy the same condition in Step 2 (namely, one of (a)-(f)), then $f(P^x_{ij}) \ne f(P^y_{ij})$ since $x_0+\lfloor x /s\rfloor \ne x_0+h+\lfloor x /s\rfloor \equiv y_0 + \lfloor y/s \rfloor \mod |i|$. If $\lfloor q/2 -1\rfloor s\le x < \lfloor q/2\rfloor s$ and $y > n - j$, then $x_0=y_0 > s-j$ ($j< \lfloor s/2\rfloor$) and $y \ge (q-1) s$, and so $\lfloor y/s \rfloor - \lfloor x/s \rfloor \ge \lceil q/2\rceil \ge |i| > h$, which contradicts $y = x+h s$. For other ranges of $x$ and $y$, we have $x_0=y_0$ and so $f(P^x_{ij}) \ne f(P^y_{ij})$.
\medskip
\textsf{Subcase 2.4:}~$x = y+hs-n$. In this case we have $x < h s$ and $y> n- h s = (q-h)s +r $. Since $h< |i| \le \lceil q/2\rceil$, we then have $0 \le x < \lceil (q - 2)/2 \rceil s$ and $\lfloor (q+2)/2\rfloor s \le y < n$. Also we have $y_0=x_0+ r\mod{s}$ and $\lfloor y/s \rfloor = \lfloor (x+r)/s \rfloor + q - h$.
We have the following: (i) when $r=0$ (which implies $y_0 = x_0$) or $n-j < y$, $f(P^x_{ij})_1$ and $f(P^y_{ij})_1$ differ by at least $i$ by Step 2;
(ii) when $x_0\le s-j$ and $y_0 > s-j$, $f(P^x_{ij})_1 \equiv x_0+\lfloor x/s \rfloor\mod |i|$ and $f(P^y_{ij})_1=(x_0 + r) +(\lfloor x/s \rfloor + q - h) -q -r \equiv x_0+\lfloor x / s \rfloor- h \mod |i|$, and so $f(P^x_{ij})_1 \ne f(P^y_{ij})_1$;
(iii) when $x_0> s - j$ and $y_0 \le s - j$, we have $y_0 = x_0 + r - s$ and $\lfloor y/s \rfloor = \lfloor x/s \rfloor + 1+ q - h$; hence $f(P^x_{ij})_1 \equiv i + ( x_0 + \lfloor x/s \rfloor\mod |i|)$ and $f(P^y_{ij})_1 \equiv i+((x_0 + r - s ) +( \lfloor x /s \rfloor+ 1+ q - h)+s-q-r-1\mod |i|) \equiv i+(x_0+\lfloor x /s \rfloor - h \mod |i|)$, implying $f(P^x_{ij})_1 \ne f(P^y_{ij})_1$;
(iv) when $x_0 > s - j$ and $y_0 > s - j$, or $x_0\le s-j$ and $y_0 \le s-j$, $f(P^x_{ij})_1$ and $f(P^y_{ij})_1$ differ by at least $i$ by Step 2.
In summary, whenever $P^x_{ij}$ and $P^y_{ij}$ share a link, they are assigned different colours.
\medskip
\textsf{Case 3:}~$j<0$. If $f(P^x_{ij})=f(P^y_{ij})$, then $f(P^x_{(-i)(-j)})=f(P^y_{(-i)(-j)})$ by Construction \ref{colouring}. Thus, by what we proved in Cases 1 and 2, $P^x_{(-i)(-j)}$ and $P^y_{(-i)(-j)}$ do not have any common link. Therefore, $P^x_{ij}$ and $P^y_{ij}$ have no common link.
So far we have proved that the colouring $f: {\cal R} \rightarrow \ZZ^3$ is arc-conflict-free.
The number of colours used by $f$ is $|f({\cal R})|$. We now estimate this number and thus obtain the required upper bounds for ${\overrightarrow{w}} (C_n(1,s),{\cal R})$ by using ${\overrightarrow{w}} (C_n(1,s),{\cal R}) \le |f({\cal R})|$. For fixed $i$ and $j$, $f$ uses $4j$ colours if $j \ge |i|$, $2|i|+j$ colours if $0 \le j < |i|$, and no new colours if $j<0$. Note that $|i| \le \lceil q/2\rceil$ and $|j| \le \lfloor s/2\rfloor$ as mentioned earlier.
Thus, if $j \ge |i|$ and $j > \lceil q/2\rceil$, then $-\lceil q/2\rceil \le i \le \lfloor q/2\rfloor$; and if $0 \le j \le |i| -1$ and $|i| -1 > \lfloor s/2\rfloor$, then $0 \le j \le \lfloor s/2\rfloor$. Therefore, by setting $\gamma_1 := \min \{\lfloor s/2\rfloor, \lceil q/2\rceil\}$ and $\gamma_2 := \min\{\lfloor s/2\rfloor + 1, \lceil q/2\rceil\}$ and noting $\sum_{i=-k}^k 2|i| =\sum_{i=1}^k 4 i $, we obtain
\begin{equation}
\label{eq:sumf}
|f({\cal R})| = \sum_{ j = 1}^{\gamma_1}\sum_{i = - j}^{j} 4j
+ \sum_{ j = 1 + \gamma_1 }^{\lfloor s/2\rfloor} \sum_{i = - \lceil q/2\rceil}^{\lceil q/2\rceil} 4j
+ \sum_{i = 1}^{\gamma_2} \sum_{j=0 }^{ i-1 } \left( 4i+2j \right)
+ \sum_{i = 1 +\gamma_2 }^{\lceil q/2\rceil} \sum_{j=0}^{\lfloor s/2\rfloor} \left( 4 i+ 2j \right).
\end{equation}
If $\lfloor s/2\rfloor \le \lceil q/2\rceil - 2$, then (\ref{eq:sumf}) is equal to
\begin{equation}
\label{eq:wb1}
\frac{1}{6}\left\lfloor \frac{s+2}{2}\right\rfloor \left( 3 q^2+ 6q + (3q+10)\left\lfloor \frac{s}{2}\right\rfloor + 8\left\lfloor \frac{s}{2}\right\rfloor^2\right) + \frac{\epsilon(q)}{6}\left\lfloor \frac{s+2}{2}\right\rfloor \left( 3\left\lfloor \frac{s}{2}\right\rfloor + q+\frac{3}{2}\right).
\end{equation}
If $\lceil q/2\rceil -1 \le \lfloor s/2\rfloor \le \lceil q/2\rceil$, then (\ref{eq:sumf}) is equal to
\begin{equation}
\label{eq:wb2}
\frac{5 q^3+12 q^2 + 4q}{24} + \frac{2}{3} \left\lfloor \frac{s}{2}\right\rfloor \left\lfloor \frac{s+2}{2}\right\rfloor \left(4 \left\lfloor \frac{s}{2}\right\rfloor + 5 \right) + \epsilon(q)\frac{ 5q^2 +13 q + 7}{8}.
\end{equation}
If $\lfloor s/2\rfloor \ge \lceil q/2\rceil +1$, then (\ref{eq:sumf}) is equal to
\begin{equation}
\label{eq:wb3}
\frac{q^3 +12 q^2 + 20 q}{24} + (2 q + 2) \left\lfloor \frac{s}{2}\right\rfloor \left\lfloor \frac{s+2}{2}\right\rfloor + \epsilon(q) \left( \frac{ q^2 + 9 q + 11 }{8} + 2 \left\lfloor \frac{s+2}{2}\right\rfloor \left\lfloor \frac{s}{2}\right\rfloor \right).
\end{equation}
Recall from (\ref{eq:ka}) that $\kappa(a)=a+(\epsilon(s)+\epsilon(q))/2$. Thus, for $s = \sqrt{n-r + (\kappa(a))^2} + \kappa(a)$, we have $\lfloor s/2\rfloor = \lceil q/2\rceil + a$ as $q=(n-r)/s$. Therefore, by applying $\lfloor s/2\rfloor \le s/2$ in (\ref{eq:wb1})-(\ref{eq:wb3}), we obtain upper bounds for $|f({\cal R})|$ which yields the upper bounds in (\ref{eq:thw1})-(\ref{eq:thw3}).
To prove that the upper bounds in (\ref{eq:thw1})-(\ref{eq:thw3}) also apply to $w(C_n(1,s),{\cal R})/2$,
we modify the definition of $f$ as follows. Define $f$ in the same as in Construction \ref{colouring} except that in Step 3 we redefine $f(P^x_{ij})=-f(P^x_{(-i)(-j)})$ for $j<0$. Obviously, $f(P^x_{i j}) \ne f(P^{y}_{k l})$ if $k \ne i$ or $l \ne j$. Moreover, when $P^x_{i j}$ and $P^{y}_{i j}$ share an edge, they share an arc and so are assigned distinct colours by the discussion above. Therefore, this modified colouring $f$ is edge-conflict-free. Since it uses twice as many colours as in the directed version, the upper bounds in (\ref{eq:thw1})-(\ref{eq:thw3}) are also upper bounds for $w(C_n(1,s),{\cal R})/2$.
\qed
\bigskip
{\bf Acknowledgements}~~The authors appreciate Professor Graham Brightwell for his comments which led to Lemma \ref{lem:ssqrtn} and subsequent improvement of a lower bound. The authors also acknowledge an anonymous referee for her/his helpful comments. Mokhtar was supported by MIFRS and MIRS of the University of Melbourne and Zhou by a Future Fellowship (FT110100629) of the Australian Research Council.
\small
\bibliographystyle{amsplain}
|
\section{Introduction}
Neighborhood frames as a generalization of Kripke semantics for modal logic were invented independently by Dana Scott \cite{Scott1970-SCOAOM-2} and Richard Montague \cite{montague1970}.
Neighborhood semantics is more general than Kripke semantics and in case of normal reflexive and transitive logics coincide with topological semantics.
In this paper we consider product of neighborhood frames, which was introduced by Sano in \cite{Sano11:AxiomHybriProduMonotNeighFrame}. It is a generalization of product of topological spaces\footnote{``Product of topological spaces'' is a well-known notion in Topology but it is different from what we use here (for details see \cite{benthem:MultimodalLogicsProductsTopologies})} presented in \cite{benthem:MultimodalLogicsProductsTopologies}.
The product of neighborhood frames is defined in the vein of the product of Kripke frames (see \cite{Sege73:Twodimodallogic} and \cite{Sheh78:Twodimodallogic}). But, there are some differences.
In any product of Kripke frames axioms of commutativity and Church-Rosser property are valid. Nonetheless,
as it was shown in \cite{benthem:MultimodalLogicsProductsTopologies}, the logic of the products of all topological spaces is the fusion of logics $\logic{S4} \otimes \logic{S4}$.
In his recent work \cite{Urid11:ModalLogicBitopRatioPlane} Uridia considers derivational semantics for products of topological spaces.
He proves that the logic of all topological spaces is the fusion of logics $\logic{D4} \otimes \logic{D4}$.
And in fact $\logic{D4} \otimes \logic{D4}$ is complete w.r.t.~the product of the rational numbers $\mathbb{Q}$.
Derivational and topological semantics can be considered as a special case of neighborhood semantics. So the result of \cite{Urid11:ModalLogicBitopRatioPlane} and corresponding result for $\logic{S4}$ from \cite{benthem:MultimodalLogicsProductsTopologies} can be obtained as corollaries from the main result of this paper.
Neighborhood frames are usually considered in the context of non-normal logics since they are usually complete w.r.t. non-normal logics and Kripke frames are not.
In this paper, however, we will consider only monotone neighborhood frames, that correspond to normal modal logics.
In some sense the results of this paper (and of \cite{benthem:MultimodalLogicsProductsTopologies}, \cite{Urid11:ModalLogicBitopRatioPlane}) shows that neighborhood semantics in some sense is more natural for products of normal modal logics, since there are no need to add extra axioms (at least in some cases).
\section{Language and logics}
In this paper we study propositional modal logic with modal operators. A formula is defined recursively as follows:
$$
\phi ::= p\; |\;\bot \; | \; \phi \to \phi \; | \; \Box_i \phi,
$$
where $p\in \mathrm{PROP}$ is a propositional letter and $\Box_i$ is a modal operator.
Other connectives are introduced as abbreviations: classical connectives are expressed through $\bot$ and $\to$,
dual modal operators $\Diamond_i$ are expressed as follows $\Diamond_i = \lnot\Box_i\lnot$.
\begin{definition} A \emph{normal logic\/} (or \emph{a logic,\/} for short) is
a set of modal formulas closed under Substitution $\left
(\frac{A(p_i)}{A(B)}\right )$, Modus Ponens $\left (\frac{A,\,
A\to B}{B}\right )$ and two Generalization rules $\left
(\frac{A}{\Box_i A}\right)$; containing all
classic tautologies and the following axioms
$$
\begin {array}{l}
\Box_i (p\to q)\to (\Box_i p\to \Box_i q).
\end{array}
$$
$\logic{K_n}$ denotes \emph{the minimal normal modal logic with $n$ modalities} and $\logic{K} = \logic{K_1}$.
\end{definition}
Let $\logic{L}$ be a logic and let $\Gamma$ be a set of formulas, then $\logic{L} +
\Gamma$ denotes the minimal logic containing $\logic{L}$ and $\Gamma$. If
$\Gamma = \set{A}$, then we write $\logic{L} + A$ rather than $\logic{L}+ \{ A \}$
\begin{definition}
Let $\logic{L_1}$ and $\logic{L_2}$ be two modal logics with one modality $\Box$ then \emph{fusion} of these logics is
\[
\logic{L_1} \otimes \logic{L_2} = K_2 + \logic{L_1}_{(\Box \to \Box_1)} + \logic{L_2}_{(\Box \to \Box_2)};
\]
where $\logic{L_i}_{(\Box \to \Box_i)}$ is the set of all formulas from $\logic{L_i}$ where all $\Box$ replaced by $\Box_i$.
\end{definition}
In this paper we consider the following four well-known logics:
\begin{align*}
\logic{D} &= \logic{K} + \Box p \to \diamondsuit p;\\
\logic{T} &= \logic{K} + \Box p \to p;\\
\logic{D4} &= \logic{D} + \Box p \to \Box\Box p;\\
\logic{S4} &= \logic{T} + \Box p \to \Box\Box p.\\
\end{align*}
\section{Kripke frames}
The notion of Kripke frames and Kripke models is well known (see \cite{blackburn_modal_2002}), so we only define special kind of frames that we are using in this paper. We can call them \emph{fractal} frames because their basic property is that any cone is isomorphic to the whole frame. In particular, we consider four types of infinite trees with fixed branching: irreflexive and transitive, reflexive and transitive, irreflexive and non-transitive (any point sees only next level) and reflexive and non-transitive.
\begin{definition}
Let $A$ be a nonempty set.
\[
A^* = \setdef[ a_1 \ldots a_k]{ a_i \in A}
\]
be the set of all finite sequences of elements from $A$, including the empty sequence $\Lambda$. Elements from $A^*$ we will denote by letters with an arrow (e.g. $\ve{a} \in A^*$) The length of sequence $\ve{a} = a_1 \ldots a_k$ is $k$ (Notation: $l(\ve{a}) = k$) the length of the empty sequence equals 0 ($l (\Lambda) = 0$). Concatenation is denoted by ``$\cdot$'': $(a_1\ldots a_k) \cdot (b_1 \ldots b_l) = \ve{a}\cdot \ve{b} = a_1\ldots a_k b_1 \ldots b_l$.
\end{definition}
\begin{definition}
Let $A$ be a nonempty set. We define an infinite frame $F_{in}[A] = (A^*, R)$, such that for $\ve{a}, \ve{b} \in A^*$
\[
\ve{a} R \ve{b} \iff \exists x\in A \left( \ve{b} = \ve{a}\cdot x \right).
\]
We also defined
\begin{align*}
F_{rn}[A] &= (A^*, R^r), \hbox{ where $R^r = R \cup Id$ --- reflexive closure};\\
F_{it}[A] &= (A^*, R^*), \hbox{ where $R^* = \bigcup\limits_{i=1}^\infty R^i$ --- transitive closure};\\
F_{rt}[A] &= (A^*, R^{r*}).
\end{align*}
So ``$t$'' stands for transitive, ``$n$'' --- for non-transitive ``$r$'' for reflexive and ``$i$'' for irreflexive.
\end{definition}
The following easy-to-prove proposition shows that frames $F_{\xi \eta}[A]$ (where $\xi \in \set{i, r}$ and $\eta \in \set{t, n})$) are indeed fractal.
\begin{proposition}
Let $F = F_{\xi \eta}[A] = (A^*, R)$ then
\[
\ve{a} R (\ve{a}\cdot\ve{c}) \iff \Lambda R \ve{c}.
\]
\end{proposition}
\begin{definition}\label{def:F_fusion}
Let $F_1 = F_{\xi_1 \eta_1}[A] = (A^*, R_1)$ be and $F_2 = F_{\xi_2 \eta_2}[B] = (B^*, R_2)$, where $\xi_1, \xi_2 \in \set{i, r}$ and $\eta_1, \eta_2, \in \set{t, n})$, $A \cap B = \varnothing$, $A = \set{a_1, a_2, \ldots}$ and $B = \set{b_1, b_2, \ldots }$ then we define frame $F_1 \otimes F_2 = (W, R'_1, R'_2)$, as follows
\begin{align*}
W &= (A \sqcup B)^* \\
\ve{x} R'_1 \ve{y} &\iff \ve{y} = \ve{x}\cdot \ve{z} \hbox{ for some } \ve{z} \in A^* \hbox{ such that } \Lambda R_1 \ve{z} \\
\ve{x} R'_2 \ve{y} &\iff \ve{y} = \ve{x}\cdot \ve{z} \hbox{ for some } \ve{z} \in B^* \hbox{ such that } \Lambda R_2 \ve{z}
\end{align*}
\end{definition}
\begin{proposition}[\cite{KracWolt91:Propeindepaxiombimodlogic}, \cite{FinSch96}]
Let $F_1$ and $F_2$ be as in Definition \ref{def:F_fusion} then
\begin{equation}
\mathop{Log}(F_1 \otimes F_2) = \mathop{Log}(F_1) \otimes \mathop{Log}(F_2).
\end{equation}
\end{proposition}
Let as define four frames: $F_{in}=F_{in}[\omega]$, $F_{rn} = F_{rn}[\omega]$, $F_{it} = F_{it}[\omega]$ and $F_{rt} = F_{rt}[\omega]$
\begin{proposition}\label{prop:compl_Kframes}
For just defined frames
\begin{enumerate}
\item $\mathop{Log}(F_{in}) = \logic{D}$;
\item $\mathop{Log}(F_{rn}) = \logic{T}$;
\item $\mathop{Log}(F_{it}) = \logic{D4}$;
\item $\mathop{Log}(F_{rt}) = \logic{S4}$.
\end{enumerate}
\end{proposition}
\section{Neighborhood frames}
In this section we consider neighborhood frames. All definitions and lemmas of this section are well-known and can be found in \cite{Segerberg1971} and \cite{Chellas1980}.
\begin{definition}
A \emph{(monotone) neighborhood frame} (or an n-frame) is a pair $\mathfrak X = (X, \tau)$, where $X$ is a nonempty set and $\tau: X \to 2^{2^X}$ such that $\tau(x)$ is a filter on $X$ for any $x$. We call function $\tau$ the \emph{neighborhood function} of $\mathfrak X$ and sets from $\tau(x)$ we call \emph{neighborhoods of $x$}.
\emph{The neighborhood model} (n-model) is a pair $(\mathfrak X, V)$, where $\mathfrak X = (X, \tau)$ is a n-frame and $V: PV \to 2^X$ is a \emph{valuation}.
In a similar way we define \emph{neighborhood 2-frame} (n-2-frame) as $(X, \tau_1, \tau_2)$ such that $\tau_i (x)$ is a filter on $X$ for any $x$, and a \emph{n-2-model}.
\end{definition}
\begin{definition}
\emph{The valuation of a formula} $\varphi$ at a point of a n-model $M = (\mathfrak X, V)$ is defined by induction as usual for boolean connectives and for modalities as follows
\[
M, x \models \Box_i \psi \iff \exists V\in \tau_i(x) \forall y\in V (M,y \models \psi).
\]
Formula is valid in a n-model $M$ if it is valid at all points of $M$ (notation $M \models \varphi$).
Formula is valid in a n-frame $\mathfrak X$ if it is valid in all models based on $\mathfrak X$ (notation $\mathfrak X \models \varphi$).
We write $\mathfrak X \models L$ if for any $\varphi \in L$, $\mathfrak X\models \varphi$.
Logic of a class of n-frames $\mathcal{C}$ as $\mathop{Log}(\mathcal{C}) = \setdef[\varphi]{\mathfrak X \models \varphi \hbox{ for some } \mathfrak X \in \mathcal{C}}$.
For logic $L$ we also define $nV(L) =\setdef[\mathfrak X]{\hbox{$\mathfrak X$ is an n-frame and } \mathfrak X\models L}$.
\end{definition}
\begin{definition}
Let $F = (W, R)$ be a Kripke frame. We define n-frame $\mathcal{N}(F) = (W, \tau)$ as follows. For any $w \in W$
\[
\tau (w) = \setdef[U]{ R(w) \subseteq U \subseteq W}.
\]
\end{definition}
\begin{lemma}\label{lem:n-frame_from_kframe}
Let $F = (W, R)$ be a Kripke frame. Then
\[
\mathop{Log}(\mathcal{N}(F)) = \mathop{Log}(F).
\]
\end{lemma}
The proof is straightforward.
\begin{definition} \label{def:bounded_morphism}
Let $\mathfrak X = (X, \tau_1, \ldots)$ and $\mathcal Y = (Y, \sigma_1, \ldots)$ be n-frames. Then function $f: X \to Y$ is a \emph{bounded morphism} if
\begin{enumerate}
\item $f$ is surjective;
\item for any $x\in X$ and $U \in \tau_i(x)$ $f(U) \in \sigma_i(f(x))$;
\item for any $x\in X$ and $V \in \sigma_i(f(x))$ there exists $U \in \tau_i(x)$, such that $f(U) \subseteq V$.
\end{enumerate}
In notation $f: \mathfrak X \twoheadrightarrow \mathcal Y$.
\end{definition}
\begin{lemma}\label{lem:pmorhism4n-frame}
Let $\mathfrak X = (X, \tau_1, \ldots)$, $\mathcal Y = (Y, \sigma_1, \ldots)$ be n-frames, $f: \mathfrak X \twoheadrightarrow \mathcal Y$ $V'$ is a valuation on $\mathcal Y$. We define $V(p) = f^{-1}(V'(p))$. Then
\[
\mathfrak X, V, x \models \varphi \iff \mathcal Y, V', f(x) \models \varphi.
\]
\end{lemma}
The proof is by standard induction on length of formula.
\begin{corollary}
If $f: \mathfrak X \twoheadrightarrow \mathcal Y$ then $\mathop{Log}(\mathcal Y) \subseteq \mathop{Log}(\mathfrak X)$.
\end{corollary}
\begin{definition}
Let $\mathfrak X_1=(X_1, \tau_1)$ and $\mathfrak X_2 =(X_2, \tau_2)$ be two n-frames. Then \emph{the product} of these n-frames is an n-2-frame defined as follows
\[
\begin{array}{l}
\mathfrak X_1 \times \mathfrak X_2 = (X_1 \times X_2, \tau_1', \tau_2'),\\
\tau_1'(x_1, x_2) = \setdef[U\subseteq X_1 \times X_2]{\exists V( V \in \tau_1(x_1) \;\&\; V \times \set{x_2} \subseteq U)},\\
\tau_2'(x_1, x_2) = \setdef[U\subseteq X_1 \times X_2]{\exists V( V \in \tau_2(x_2) \;\&\; \set{ x_1} \times V \subseteq U)}.
\end{array}
\]
\end{definition}
\begin{definition}
For two unimodal logics $\logic{L_1}$ and $\logic{L_2}$ we define \emph{n-product} of them as follows
\[
\logic{L_1} \times_n \logic{L_2} = \mathop{Log} (\setdef[\mathfrak X_1 \times \mathfrak X_2]{\mathfrak X_1\in nV(L_1) \;\&\; \mathfrak X_2 \in nV(L_2)})
\]
\end{definition}
Note that $\mathfrak X_1 \times \mathfrak X_2$ if we forget about one of its neighborhood functions say $\tau'_2$ then $\mathfrak X_1 \times \mathfrak X_2$ will be a disjoint union of $\logic{L_1}$ n-frames. Hence
\begin{proposition}[\cite{Sano11:AxiomHybriProduMonotNeighFrame}]\label{prop:fusin_of_n-frames}
For two unimodal logics $\logic{L_1}$ and $\logic{L_2}$
\[
\logic{L_1} \otimes \logic{L_2} \subseteq \logic{L_1} \times_n \logic{L_2}.
\]
\end{proposition}
\section{Main construction}
The construction in this section was inspired by \cite{benthem:MultimodalLogicsProductsTopologies}, but it is not a straightforward generalization.
In case of $\logic{S4 \times_n S4}$ it is, in essence, very similar to the construction in \cite{benthem:MultimodalLogicsProductsTopologies}.
However, here we operate only with words (finite or infinite), and not with numbers and fractions. It makes proofs shorter and allows us to generalize the results to non-transitive cases.
\newcommand{\mathop{st}}{\mathop{st}}
Let $F = (A^*, R) = F_{\xi \eta}[A]$ and $0 \notin A$. We define set of ``pseudo-infinite'' sequences
\[
X = \setdef[a_1 a_2 \ldots]{a_i \in A \cup \set{0}\ \&\ \exists N \forall k\ge N (a_k = 0)}.
\]
Define $f_F:X \to A^*$ which ``fogets'' all zeros. For $\alpha \in X$ such that $\alpha = a_1 a_2 \ldots$ we define
\begin{align*}
\mathop{st}(\alpha) &= \min \setdef[N]{\forall k\ge N (a_k = 0)};\\
\alpha|_k &= a_1 \ldots a_k;\\
U_k(\alpha) &= \setdef[\beta \in X]{\alpha |_m = \beta |_m \ \&\ f_F(\alpha) R f_F(\beta), \hbox{ where } m = \max(k, \mathop{st}(\alpha))}.
\end{align*}
\begin{lemma}\label{lem:U_m-base}
$U_k (\alpha) \subseteq U_m (\alpha)$ whenever $k \ge m$.
\end{lemma}
\begin{proof}
Let $\beta \in U_k (\alpha)$. Since $\alpha |_k = \beta |_k$ and $k \ge m$ then $\alpha |_m = \beta |_m$. Hence, $\beta \in U_m (\alpha)$.
\end{proof}
\begin{definition}\label{def:n-frame_from_fframe}
Due to Lemma \ref{lem:U_m-base} sets $U_n(\alpha)$ forms a filter base. So we can define
\begin{align*}
\tau (\alpha) &- \hbox{the filter with base } \setdef[U_n(\alpha)]{n \in \omega};\\
\mathcal{N_\omega}(F) &= (X,\tau) \hbox{ --- is \emph{the n-frame based on} $F$.}
\end{align*}
\end{definition}
\begin{lemma}\label{lem:bmorphism_wframe2nframe}
Let $F = (A^*, R) = F_{\xi, \eta}[A]$ then
\[
f_F : \mathcal{N_\omega}(F) \twoheadrightarrow \mathcal{N}(F).
\]
\end{lemma}
\begin{proof}
From now on in this proof we will omit the subindex in $f_F$.
Let $\mathcal{N_\omega}(F) = (X, \tau)$.
Since for any $\ve{x} \in A^*$ sequence $\ve{x} \cdot 0^\omega \in X$ and $f(\ve{x} \cdot 0^\omega) = \ve{x}$ then $f$ is surjective.
Assume, that $x\in X$ and $U \in \tau(x)$. We need to prove that $R(f(x)) \subseteq f(U)$.
There is $m$ such that $U_m(x) \subseteq U$ and since $f(U_m(x)) = R(f(x))$ then
\[
R(f(x)) = f(U_m(x)) \subseteq f(U).
\]
Assume that $x\in X$ and $V$ is a neighborhood of $x$, i.e.\/ $R(f(x)) \subseteq V$. We need to prove that there exists $U \in \tau(x)$, such that $f(U) \subseteq V$. As $U$ we take $U_m(x)$ for some $m \ge st(x)$, then
\[
f(U_m(x))=R(f(x)) \subseteq V.
\]
\end{proof}
\begin{corollary}\label{cor:logicNf}
For frame $F = F_{\xi \eta}[A]$ $\mathop{Log}(\mathcal{N_\omega}(F)) \subseteq \mathop{Log}(F)$.
\end{corollary}
\begin{proof}
It follows from Lemmas \ref{lem:n-frame_from_kframe}, \ref{def:bounded_morphism} and \ref{lem:bmorphism_wframe2nframe}
\[
\mathop{Log}(\mathcal{N_\omega}(F)) \subseteq \mathop{Log}(\mathcal{N}(F)) = \mathop{Log}(F).
\]
\end{proof}
\begin{proposition}\label{prop:logicsOf_omega_Nf_framas}
Let $F_{in} = F_{in}[\omega]$, $F_{rn} = F_{rn}[\omega]$, $F_{it} = F_{it}[\omega]$ and $F_{rt} = F_{rt}[\omega]$ then
\begin{enumerate}
\item\label{it1} $\mathop{Log}(\mathcal{N_\omega}(F_{in})) = \logic{D}$;
\item\label{it2} $\mathop{Log}(\mathcal{N_\omega}(F_{rn})) = \logic{T}$;
\item\label{it3} $\mathop{Log}(\mathcal{N_\omega}(F_{it})) = \logic{D4}$;
\item\label{it4} $\mathop{Log}(\mathcal{N_\omega}(F_{rt})) = \logic{S4}$.
\end{enumerate}
\end{proposition}
\begin{proof}
In all these cases the inclusion from left to right is covered by Corollary \ref{cor:logicNf} and Proposition \ref{prop:compl_Kframes}.
Let us check the inclusion in converse direction.
(\ref{it1}). It is easy to check that n-frame $\mathfrak X = (X, \tau) \models \logic{D}$ iff for each $x \in X$ $\varnothing \notin \tau(x)$. For $\mathcal{N_\omega}(F_{in})$ and $\mathcal{N_\omega}(F_{it})$ it obviously true.
(\ref{it2}). It is easy to check that n-frame $\mathfrak X = (X, \tau) \models \logic{T}$ iff $x \in U \in \tau(x)$ for each $x$ and $U$. For $\mathcal{N_\omega}(F_{rn})$ and $\mathcal{N_\omega}(F_{rt})$ it is obviously true.
(\ref{it3}) and (\ref{it4}). It is well-known (see e.g.{} \cite{Hans03:Monotmodallogic}) that $\mathfrak X = (X, \tau) \models \Box p \to \Box \Box p$ iff for each $U \in \tau(x)$ $\setdef[y]{ U \in \tau(y)} \in \tau (x)$.
Indeed, it follows from the fact that for any $y \in U_m (x)$ and any $k$ $U_k(y) \subseteq U_m(x)$.
\end{proof}
Let $F_1 = (A^*, R_1) = F_{\xi_1\eta_1}[A]$ and $F_2 = (B^*, R_2) = F_{\xi_2\eta_2}[B]$ we assume that $A \cap B = \varnothing$, $A = \set{a_1, a_2, \ldots}$ and $B = \set{b_1, b_2, \ldots }$. Consider the product of n-frames $\mathfrak X_1 = (X_1, \tau_1) = \mathcal{N_\omega}(F_1)$ and $\mathfrak X_2 = (X_2, \tau_2) = \mathcal{N_\omega}(F_2)$
\[
\mathfrak X = (X_1 \times X_2, \tau_1', \tau_2') = \mathcal{N_\omega}(F_1) \times_n \mathcal{N_\omega}(F_2).
\]
We define function $g: \mathfrak X_1 \times \mathfrak X_2 \to (A \cup B)^*$ as follows. For $(\alpha, \beta) \in \mathfrak X_1 \times \mathfrak X_2$, such that $\alpha = x_1 x_2 \ldots$ and $\beta = y_1 y_2 \ldots$, $x_i \in A \cup \set{0}$, $y_j \in B \cup \set{0}$, we define
$g(\alpha, \beta)$ to be the finite sequence which we get after eliminating all zeros from the infinite sequence
$x_1y_1x_2y_2 \ldots$.
\begin{lemma}\label{lem:main}
Function $g$ defined above is a bounded morphism: $g: \mathfrak X \twoheadrightarrow \mathcal{N}(F_1 \otimes F_2)$.
\end{lemma}
\begin{proof}
Let $\ve{z} = z_1 z_2 \ldots z_n \in (A \cup B)^*$. Define for $i\le n$
\[
x_i = \left\{
\begin{array}{l}
z_i, \hbox{ if $z_i \in A$};\\
0, \hbox{ if $z_i \notin A$};
\end{array}
\right.\qquad
y_i = \left\{
\begin{array}{l}
z_i, \hbox{ if $z_i \in B$};\\
0, \hbox{ if $z_i \notin B$}.
\end{array}
\right.
\]
Let $\alpha = x_1 x_2 \ldots x_n 0^\omega$ and $\beta = y_1 y_2 \ldots y_n 0^\omega$ then $g(\alpha, \beta) = \ve{z}$.
Hence $g$ is surjective.
The next two conditions we check only for $\tau_1$ and for $\tau_2$ it is similar.
Assume, that $(\alpha, \beta) \in X_1 \times X_2$ and $U \in \tau_1(\alpha, \beta)$. We need to prove that $R_1'(g(\alpha, \beta)) \subseteq g(U)$.
There is $m > \max\set{st(\alpha), st(\beta)}$ such that $U'_m(\alpha) \times \set{\beta} \subseteq U$ and since $g(U'_m(\alpha) \times \set{\beta}) = R_1'(g(\alpha, \beta))$ then
\[
R_1'(g(\alpha, \beta)) = g(U'_m(\alpha)\times \set{\beta}) \subseteq g(U);
\]
where $U'_m(\alpha)$ is the corresponding neighborhood from $\mathfrak X_1$.
Assume that $(\alpha, \beta) \in X_1 \times X_2$ and $R_1'(g(\alpha, \beta)) \subseteq V$. We need to prove that there exists $U \in \tau'_1(\alpha, \beta)$, such that $g(U) \subseteq V$. As $U$ we take $U'_m(\alpha)\times \set{\beta}$ for some $m > \max\set{st(\alpha), st(\beta)}$, then
\[
g(U_m' (\alpha) \times \set{\beta}) = R_1'(g(\alpha, \beta)) \subseteq V.
\]
\end{proof}
\begin{corollary}\label{cor:logicNftimesNf}
Let $F_1 = (A^*, R_1) = F_{\xi_1\eta_1}[A]$ and $F_2 = (B^*, R_2) = F_{\xi_2\eta_2}[B]$ then
$\mathop{Log}(\mathcal{N_\omega}(F_1) \times_n \mathcal{N_\omega}(F_2)) \subseteq \mathop{Log}(F_1) \otimes \mathop{Log}(F_2)$.
\end{corollary}
It immediately follows from Lemmas \ref{lem:main}, \ref{lem:pmorhism4n-frame} and Proposition \ref{prop:fusin_of_n-frames}.
\begin{corollary}\label{cor:main}
Let $F_1, F_2 \in \set{F_{in}, F_{rn}, F_{it}, F_{rt}}$ then
$\mathop{Log}(\mathcal{N_\omega}(F_1) \times_n \mathcal{N_\omega}(F_2)) = \mathop{Log}(F_1) \otimes \mathop{Log}(F_2)$.
\end{corollary}
\begin{proof}
The left-to-right inclusion follows from Corollary \ref{cor:logicNftimesNf}.
To prove right-to-left inclusion we notice that due to Proposition \ref{prop:logicsOf_omega_Nf_framas} $\mathop{Log}(\mathcal{N_\omega}(F_i)) = \mathop{Log}(F_i)$ ($i=1,\,2$) and due to Proposition \ref{prop:fusin_of_n-frames}
\[
\mathop{Log}(\mathcal{N_\omega}(F_1)) \otimes \mathop{Log}(\mathcal{N_\omega}(F_2)) \subseteq \mathop{Log}(\mathcal{N_\omega}(F_1)) \times_n \mathop{Log}(\mathcal{N_\omega}(F_2)).
\]
\end{proof}
\section{Completeness results}
\begin{theorem}
Let $\logic{L_1}, \logic{L_2} \in \set{\logic{S4}, \logic{D4}, \logic{D}, \logic{T}}$ then
\[
\logic{L_1} \times_n \logic{L_2} = \logic{L_1} \otimes \logic{L_2}.
\]
\end{theorem}
\begin{proof}
Logics $\logic{L_1} = \mathop{Log}(F_1)$ and $\logic{L_2}= \mathop{Log}(F_2)$ for some $F_1, F_2 \in \set{F_{in}, F_{rn}, F_{it}, F_{rt}}$.
By Corollary \ref{cor:main}
\[
\logic{L_1} \times_n \logic{L_2} = \mathop{Log}(\mathcal{N_\omega}(F_1) \times_n \mathcal{N_\omega}(F_2)) = \mathop{Log}(F_1) \otimes \mathop{Log}(F_2) = \logic{L_1} \otimes \logic{L_2}.
\]
\end{proof}
The following fact was proved in \cite{benthem:MultimodalLogicsProductsTopologies}.
\begin{corollary}\label{cor:S4timesS4}
Let $\mathfrak X = (\mathbb{Q}, \tau)$ where $\mathbb{Q}$ is the set of rational numbers and $\tau$ is based on the standard topology on $\mathbb{Q}$, i.e.\/ $\tau(x) = \setdef[U]{\exists V (\hbox{$x \in V$ is open and } V \subseteq U)}$. Then
\[
\mathop{Log}(\mathfrak X \times_n \mathfrak X) = \logic{S4} \otimes \logic{S4}.
\]
\end{corollary}
\begin{proof}
Let $\mathcal{N_\omega}(F_{rt}) = (X, \tau)$. We can assume that $X = \setdef[\ve{x}\cdot 0^\omega]{\ve{x}\in \mathbb{Z}}$.
Note that $X$ is a countable set and neighborhood function $\tau$ is based on topology generated by the lexicographical order $<_l$ on $X$. According to the classical result of Cantor, since the lexicographical order on $X$ is dense, $(X, <_l)$ isomorphic to $(\mathbb{Q}, <)$ (see ) and corresponding topological spaces are homeomorphic. Hence,
\[
\mathop{Log}(\mathfrak X \times_n \mathfrak X) = \mathop{Log}( \mathcal{N_\omega}(F_{rt}) \times \mathcal{N_\omega}(F_{rt})) = \logic{S4} \otimes \logic{S4}.
\]
\end{proof}
The following fact was announced\footnote{The talk at the conference was very detailed, but to my knowledge, the full proof has not been published yet.} in \cite{Urid11:ModalLogicBitopRatioPlane}.
\begin{corollary}\label{cor:D4timesD4}
Let $\mathfrak X = (\mathbb{Q}, \tau)$ where $\mathbb{Q}$ is the set of rational numbers and $\tau$ is based on the standard topology on $\mathbb{Q}$, i.e.\/ $\tau(x) = \setdef[U]{\exists V (\hbox{$x \in V$ is open and } V \setminus \set{x} \subseteq U)}$. Then
\[
\mathop{Log}(\mathfrak X \times_n \mathfrak X) = \logic{D4} \otimes \logic{D4}.
\]
\end{corollary}
\begin{proof}
Let $\mathfrak X = (\mathbb{Q}, \tau)$, where $\tau$ is based on derivation operator in $\mathbb{Q}$. Since $(X, <_l)$ isomorphic to $(\mathbb{Q}, <)$ then
\[
\mathop{Log}(\mathfrak X \times_n \mathfrak X) = \mathop{Log}( \mathcal{N_\omega}(F_{it}) \times \mathcal{N_\omega}(F_{it})) = \logic{D4} \otimes \logic{D4}.
\]
\end{proof}
\section{Conclusion}
There are several ways to continue research. One of them is to try and extend the technique to other logics (e.g.~$\logic{K}$).
The other way is to add the third modality which corresponds to the following neighborhood function $\tau' (x, y) = \setdef[U]{\exists V_1 \in \tau_1 (x) \,\&\, \exists V_2 \in \tau_2(y) \left( V_1 \times V_2 \subseteq U \right)}$. Similar construction was considered in \cite{benthem:MultimodalLogicsProductsTopologies} for the topological semantic.
\bibliographystyle{plain}
|
\section{Secret key rates for Gaussian protocols}
We will now derive bounds upon the secret key rate for the 8 members of the Gaussian family of CVQKD protocols. In the prepare and measure setting (P\&M) protocols correspond to Alice sending either squeezed \cite{Ralph:1999p5546,Hillery:2000p8625} or coherent states \cite{Grosshans:2003p2402} and Bob measuring with either homodyne or heterodyne \cite{Weedbrook:2004p380,Weedbrook:2006p8395} detection. Each permutation can in turn be used to have Bob try and guess Alice's encoding, called direct reconciliation \cite{Grosshans:2002p377} (DR), or Alice trying to guess Bob's measurement \cite{Grosshans:2003p2402}, called reverse reconciliation (RR). One can also consider entanglement based (EB) schemes in which Alice distributes one arm of a two-mode squeezed vacuum to Bob \cite{Reid:2000p5545}. In fact these methods can be seen as equivalent since if Alice heterodyne detects one arm of a squeezed vacuum this corresponds to sending coherent states, while a homodyne detection corresponds to squeezed state protocols \cite{Grosshans:2003p526}. In the following, we will calculate the key rates assuming all parties have chosen to measure in the $\hat{x}$ basis. In general the total key rate will be the average of the this quantity and the analogous expression with $\hat{x}$ and $\hat{p}$ labels permuted.
\subsubsection{Squeezed states and Homodyne Detection}
First we turn to the protocols where both Alice and Bob homodyne detect. Here Alice and Bob randomly choose whether they measure the $\hat{x}$ of $\hat{p}$ basis and only keep those times where they agree. We work in the asymptotic regime $n \rightarrow \infty$ where the optimal attacks are known to be the collective attacks. Considering the case where Bob makes an $\hat{x}$ measurement, we can write the secret key rate for the DR protocol as
\cite{Devetak:2005p5086,GarciaPatron:2006p381,Navascues:2006p805},
\eq{K \geq I(A:B) - \chi(x_A:E)\label{k}}
where
\eqn{I(A:B) &=&\nonumber H(x_B) - H(x_B|x_A)\\
&=& H(x_A) - H(x_A|x_B) \label{iab}}
denotes the classical mutual information between Alice and Bob, with $H(x) = \int dx \hspace{0.2cm} -p(x)\log p(x)$ being the continuous Shannon entropy of the measurement strings and
\eqn{\chi(x_A:E) = S(E) - S(E|x_A)\label{ie}}
is the Holevo bound.
Substituting (\ref{iab}) and (\ref{ie}) in (\ref{k}) and comparing with the continuous conditional von Neumann entropy
\eq{S(x_A|B) = H(x_A) + \int dx_A p(x_A) S(\rho_B^{x_A}) - S(B)\label{cond}}
we have,
\eqn{K&\geq& H(x_B) + \int dx_B\hspace{0.2cm} p(x_B)\hspace{0.2cm} S(\rho_E^{x_B}) - S(E) - H(x_B|x_A)\nonumber\\
&=& S(x_B|E) - H(x_B|x_A)}
which is what one would expect from the Devetak-Winter relations \cite{Devetak:2005p5086}.
Using the entropic style uncertainty relations allows us to bound the eavesdroppers information on the relevant observable. Substituting (\ref{uc2}) from the main text and recalling that $S(p_A|B)\leq S(p_A|p_B) = H(p_B|p_A)$ we can write,
\eqn{K &\geq &\nonumber \log 4\pi - H(x_A|x_B) - S(p_B|A)\\
&=& \log 4\pi - H(x_A|x_B) - H(p_A|p_B).\label{kh}}
Thus we have bounded the secret key by an expression that depends only upon the conditional Shannon entropies and the complementarity of the measurements, both of which are directly accessible to Alice and Bob. Furthermore one can show via a differential calculation that for any probability distribution $p(x)$ the corresponding Shannon entropy is maximised for a Gaussian distribution of the same variance. In other words, Alice and Bob can bound their secret key rate for this protocol by measuring Bob's conditional variance. Substituting the Shannon entropy for a Gaussian distribution $H_G(x_B|x_A) = \log\sqrt{2\pi eV_{x_B|x_A}}$ and $c = 4\pi$ we arrive at a final expression for the key rate of
\eqn{\label{kv}K &\geq & \log 4\pi - \log 2 e \pi \sqrt{V_{x_A|x_B}V_{p_A|p_B}} \nonumber\\
&=& \log \bk{\frac{2}{e\sqrt{V_{x_A|x_B}V_{p_A|p_B}}}}\label{kh}}
The RR expression is obtained by simply permuting the labels of Alice and Bob.
The fact that this result can be pessimistic is immediately apparent if we consider the predictions for a perfect channel. Under the security proofs presented in \cite{Devetak:2005p5086,GarciaPatron:2006p381,Navascues:2006p805} if Alice and Bob share an pure EPR state (real or effective) with variance $V = \cosh(2s)$ then the key rate is always positive provided the squeezing parameter is non-zero ($s>0$). On the other hand the above result is only positive for $V_{x_A|x_B} = V_{p_A|p_B} = V_{A|B}\leq \frac{2}{e}$ which implies a squeezing parameter of $s \geq .15$ or about 1.3 dB. This result was also calculated in \cite{agnes}, however as mentioned before the proof relies on the assumption of the applicability of the entropic uncertainty relation to infinite dimensional Hilbert spaces. Furthermore Ferenczi argues that since for coherent states the directly measured conditional variances are always greater than 1 the above procedure would never predict a positive key rate for coherent state protocols. However this conclusion comes from a mistaken application of the key rate formulae as we now demonstrate.
\subsubsection{Coherent States and Homodyne Detection}
Consider a DR coherent state protocol, which in the entanglement based picture involves Alice making a heterodyne detection upon her arm of an EPR pair. Thus she first mixes her mode with vacuum resulting in two modes $A_1$ and $A_2$ upon which she measured $\hat{x}$ and $\hat{p}$ respectively. Bob is still making a homodyne detection, randomly switching between the quadratures. We will consider the case where Bob measures $\hat{x}$, with the other case following straightforwardly. The DR key rate is then given by,
\eqn{K & \geq & S(x_{A_1}|E) - H(x_{A_1}|x_B)}
After Alice's projective measurement upon $A_2$ the state $\rho_{A_1BE}$ is pure and we can again apply the entropic uncertainty relation to write,
\eqn{K &\geq &\log 4 \pi - S(p_{A_1}|B) - H(x_{A_1}|x_B)\nonumber\\
&\geq & \log 4\pi - H(p_{A_1}|p_B) - H(x_{A_1}|x_B)}
Now this formula might pose a problem, in that we do not measure $\hat{p}$ upon mode $A_1$, but this can be circumvented if we trust the devices, specifically the beamsplitter, in Alice's station for we then have $H(p_{A_1}|p_B) = H(p_{A_2}|p_B)$ which is measured. We therefore have,
\eqn{K &\geq & \log 4\pi - \log\sqrt{2\pi e V_{p_{A_2}|p_B}} - \log\sqrt{2\pi e V_{x_{A_1}|x_B}} \nonumber \\
&=& \log\frac{2}{e\sqrt{V_{x_{A_1}|x_B}V_{p_{A_2}|p_B}}}}
For reverse reconciliation the key rate is given by,
\eqn{K &\geq & S(x_B|E) - H(x_B|x_{A_1}) \nonumber \\
&\geq & \log 4\pi - S(p_B|A) - H(x_B|x_{A_1})\nonumber\\
&\geq & \log 4\pi - H(p_B|p_A) - H(x_B|x_{A_1})\nonumber\\
&\geq & \log \frac{2}{e\sqrt{V_{p_B|p_A}V_{x_B|x_{A_1}} }}\label{kcoh}}
Once again, the quantity $V_{p_B|p_A}$ is not measured directly, but provided we trust Alice's beamsplitter we have $V_{p_B|p_A} = 2V_{p_B|p_{A_2}} - 1$ allowing us to evaluate the bound.
\subsubsection{Squeezed States and Heterodyne Detection}
These protocols are essentially mirror images coherent state homodyne schemes since, in the EB picture, we now have Alice making homodyne measurements and Bob making heterodyne measurements. We now have Alice swapping between $\hat{x}$ and $\hat{p}$ measurements while Bob splits up his mode measuring $\hat{x}$ upon $B_1$ and $\hat{p}$ upon $B_2$. The DR key rate is given by,
\eqn{K &\geq & S(x_A|E) - H(x_A|x_{B_1}) \nonumber\\
&\geq & \log 4\pi - S(p_A|B) - H(x_A|x_{B_1}) \nonumber\\
&\geq & \log 4\pi - H(p_A|p_B) - H(x_A|x_{B_1})\nonumber\\
&\geq & \log \frac{2}{e \sqrt{V_{p_A|p_B}V_{x_A|x_{B_1}}}}}
where we will need to trust the beamsplitter in Bob's station to obtain $V_{p_A|p_B} = 2V_{p_A|p_{B_2}} - 1$ from the directly measured conditional variance.
The RR key rate is given by,
\eqn{K &\geq & S(x_{B_1}|E) - H(x_{B_1}|x_A) \nonumber \\
&\geq & \log 4\pi - S(p_{B_1}|A) - H(x_{B_1}|x_{A})\nonumber\\
&\geq & \log 4\pi - H(p_{B_1}|p_A) - H(x_{B_1}|x_{A})\nonumber\\
&\geq & \log \frac{2}{e\sqrt{V_{p_{B_2}|p_A}V_{x_{B_1}|x_{A}} }}}
where we have again used the form of Bob's beamsplitter to write $V_{p_{B_2}|p_A} = V_{p_{B_1}|p_A}$.
\subsubsection{Coherent States and Heterodyne Detection}
The final protocols involve Bob making a heterodyne measurement upon coherent states, or alternatively both parties making heterodyne measurements upon an two-mode squeezed vacuum. Thus there are now four modes involved $A_1$ and $B_1$ upon which $\hat{x}$ is measured and $A_2$ and $B_2$ upon which $\hat{p}$ is measured. We can consider the $\hat{x}$ and $\hat{p}$ channels separately. Note that this is actually an underestimation of the key rate as it essentially allowing Eve to devote all her resources to estimating the $\hat{x}$ and $\hat{p}$ measurements separately whereas she must in fact estimate both simultaneously. The DR key rate for $\hat{x}$ is given by,
\eqn{K &\geq & S(x_{A_1}|E) - H(x_{A_1}|x_{B_1}) \nonumber\\
&\geq & \log 4\pi S(p_{A_1}|B) - H(x_{A_1}|x_{B_1})\nonumber\\
&\geq & \log 4\pi H(p_{A_1}|p_B) - H(x_{A_1}|x_{B_1})\nonumber\\
&\geq & \log \frac{2}{e\sqrt{V_{p_{A_1}|p_B} V_{x_{A_1}|x_{B_1}}}}}
Provided we trust the beamsplitter in Bob's station we can calculate $V_{p_{A_1}|p_B} = 2 V_{p_{A_2}|p_{B_2}}-1$.
The RR key rate is given by,
\eqn{K &\geq & S(x_{B_1}|E) - H(x_{B_1}|x_{A_1}) \nonumber\\
&\geq & \log 4\pi S(p_{B_1}|A) - H(x_{B_1}|x_{A_1})\nonumber\\
&\geq & \log 4\pi H(p_{B_1}|p_A) - H(x_{B_1}|x_{A_1})\nonumber\\
&\geq & \log \frac{2}{e\sqrt{V_{p_{B_1}|p_A} V_{x_{B_1}|x_{A_1}}}}}
Provided we trust the beamsplitter in Alice and Bob's station we can calculate $V_{p_{B_1}|p_A} = 2 V_{p_{B_2}|p_{A_2}}-1$.
\section{one-sided device independence for gaussian protocols}
Another important benefit of utilising entropic uncertainty relations in QKD proofs is that they lend themselves towards one-sided device independent (1sDI) protocols \cite{Branciard:2012p8266}. For 1sDI-QKD protocols only one of Alice or Bob is untrusted and the security is now related to the recently classified steering inequalities \cite{Wiseman:2007p2187}. The 1sDI nature of entropic proofs is manifest in expressions like (\ref{kh}) in that one can immediately see it depends only upon measuring a known observable upon one side. More precisely, for a term like $V_{x_A|x_B}$ we must be sure that the observable upon Alice's side is indeed an $\hat{x}$ measurement but we need make no assumptions about what Bob is actually measuring, we simply evaluate the conditional variance. We note that the proof provided in \cite{Furrer:2012p8365} is also 1sDI, being independent of Alice's devices.
However this device independence does not necessarily extend to the protocols involving heterodyne detection or coherent states (recall that this is alternatively a heterodyne detection on Alice's side in the entanglement based picture). This is essentially because the proof to derive rates such as (\ref{kcoh}) depends upon characterising the beamsplitter on the supposedly untrusted side thereby invalidating the device independence. Nonetheless the remaining protocols, with the heterodyne detection taking place in the trusted station, are still implementable with high efficiency sources and detection opening the way to long range 1sDI-CVQKD protocols with current technology. This means that for real EPR states both DR and RR are secure provided all parties are homodyning, while Bob may safely heterodyne for an RR protocol and Alice may heterodyne for a DR protocol. Finally, for DR protocols where Alice, who controls the source, is trusted we may also safely make the equivalence between P\&M and EB schemes. Remarkably, this means that for direct reconciliation is is possible to generate 1sDI key using only coherent states. We summarise which of the 16 possible protocols are 1sDI in Table 1 in the main text.
\end{document} |
\section{Introduction}\label{intro}
Given an elliptic curve $E$ over the prime finite field $\mathbb F_p$, we let $E(\mathbb F_p)$ denote its set of $\mathbb F_p$ points.
It is well-known that $E(\mathbb F_p)$ admits the structure of an abelian group, and in fact,
\[
E(\mathbb F_p)\cong G_{m,k}:= \mathbb Z/m\mathbb Z\times\mathbb Z/mk\mathbb Z
\]
for some positive integers $m$ and $k$. It is natural to wonder which groups of the form $G_{m,k}$ arise in this way and how often they occur as $p$ varies over all primes and $E$ varies over all elliptic curves over $\mathbb F_p$.
The former problem, of characterizing which groups are realized in this way was studied in~\cite{BPS:2012, CDKS1}, while the frequency of occurrence was studied by the second and fourth named authors in~\cite{DS-MEG}.
In the present work, we explore the frequency of occurrence further.
Given a group $G$ of the form $G_{m,k}=\mathbb Z/m\mathbb Z\times\mathbb Z/mk\mathbb Z$, we set $N=|G|=m^2k$ and let $M_p(G)$ denote the weighted number of isomorphism classes of elliptic curves over $\mathbb F_p$ with group isomorphic to $G$, that is to say
\[
M_p(G)=\sum_{\substack{E/\mathbb F_p\\ E(\mathbb F_p)\cong G}}\frac{1}{|\Aut_p(E)|},
\]
where the sum is taken over all isomorphism classes of elliptic curves over $\mathbb F_p$ and
$|\Aut_p(E)|$ is the number of $\mathbb F_p$-automorphisms of $E$.
It is worth noting here that $|\Aut_p(E)|=2$ for all but a bounded number of isomorphism classes $E$ over $\mathbb F_p$, and hence
\[
M_p(G) = \frac{1}{2} \#\{ E/\mathbb F_p : E(\mathbb F_p)\cong G \} +O(1),
\]
In \cite{DS-MEG}, the authors studied the weighted number
of isomorphism classes of elliptic curves over any prime finite field with group of points isomorphic to $G$, i.e., they studied
\[
M(G):=\sum_pM_p(G) .
\]
The primes counted by $M(G)$ must lie in a very short interval near $N=|G|$.
This is because the Hasse bound implies that $p+1-2\sqrt p<N<p+1+2\sqrt p$, which is equivalent to saying that
\[
N^{-} := N+1-2\sqrt{N} < p < N+1+2 \sqrt{N}=:N^+.
\]
Even the Riemann hypothesis does not guarantee the existence of a prime in such a short interval.
Hence the main theorem of \cite{DS-MEG} can only be proven under an appropriate conjecture concerning the distribution of primes in short intervals.
In the statement below, we refer to the conjecture assumed in~\cite{DS-MEG} as the
Barban-Davenport-Halberstam (BDH) estimate for short intervals.
Before stating the main theorem of~\cite{DS-MEG}, we fix some more notation.
Given a group $G=G_{m,k}$, we let $\Aut(G)$ denote its automorphism group (as a group).
This should not be confused with $\Aut_p(E)$ as defined above,
which refers to the set of $\mathbb F_p$-automorphisms of the elliptic curve $E$.
We also define the function
\eq{define K(G)}{
K(G)=
\prod_{\ell\nmid N}\left(1-\frac{\leg{N-1}{\ell}^2\ell+1}{(\ell-1)^2(\ell+1)}\right)
\prod_{\ell\mid m}\left(1-\frac{1}{\ell^2}\right)
\prod_{\substack{\ell\mid k\\ \ell\nmid m}}\left(1-\frac{1}{\ell(\ell-1)}\right),
}
where the products are taken over all primes $\ell$ satisfying the stated conditions and $\leg{\cdot}{\ell}$ denotes the usual Kronecker symbol.
In~\cite{DS-MEG}, the function $K(G)$ was only computed for odd order groups, and its definition contained a mistake.
It was corrected to the form that we give here in~\cite{DS-MEG-corr}.
Note that the function $K(G)$ is bounded between two constants independently of the the parameters $m$ and $k$. In paraphrased form, the main theorem of~\cite{DS-MEG} is as follows.
\begin{thm}[David-Smith] \label{meg-rephrased}
Assume that the BDH estimate for short intervals holds.
Fix $A,B>0$.
Then for every nontrivial, odd order group $G=G_{m,k}$, we have that
\[
M(G)=\left(K(G) + O_{A,B}\left( \frac{1}{(\log|G|)^{B}} \right)\right)\frac{|G|^2}{|\Aut (G)|\log |G|}
\asymp \frac{mk^2}{\phi(m)\phi(k)\log k},
\]
provided that $m\le (\log k)^A$.
\end{thm}
For precise details concerning the conjecture assumed to prove Theorem~\ref{meg-rephrased},
we refer the reader to \cite{DS-MEG}. We note that the result of Theorem~\ref{meg-rephrased} is restricted to the range $m\le (\log k)^A$. However, we believe that it should hold in the range $m\le k^A$.
Proving such a result at the present time would however require an even stronger hypothesis than the one assumed in \cite{DS-MEG}. Unconditionally, it is possible to obtain upper bounds of the correct order of magnitude in this larger range.
This is the context of our first theorem.
\begin{thm} \label{CDKS} Fix $A>0$ and consider integers $m$ and $k$ with $1\le m\le k^A$. Let $G=G_{m,k}$, $N=|G|=m^2k$, and
\[
\delta = \frac{1}{N/(\phi(m)\log(2N))}
\sum_{\substack{ N^-<p\le N^+ \\ p\equiv 1\pmod m }} \sqrt{(p-N^-)(N^+-p)},
\]
and note that $\delta\ll1$ by the Brun-Titchmarsch inequality. For any fixed $\lambda>1$,
\[
\delta^\lambda \cdot \frac{|G|^2}{ |\Aut (G)|\log(2|G|)} \ll M(G) \ll \delta^{1/\lambda} \cdot \frac{|G|^2}{ |\Aut (G)|\log(2|G|)},
\]
the implied constants depending at most on $A$ and $\lambda$.
\end{thm}
Employing the above result together with the Bombieri-Vinogradov theorem, we also show that the lower bound implicit in Theorem~\ref{meg-rephrased} holds for a positive proportion of groups $G$.
\begin{thm} \label{CDKS2} Consider numbers $x$ and $y$ with $1\le x\le \sqrt{y}$.
Then there are absolute positive constants $c_1$ and $c_2$ such that
\[
M(G_{m, k}) \ge c_1\cdot \frac{|G_{m,k}|^2}{|\Aut (G_{m,k})|\log(2|G_{m,k}|)}
\]
for at least $c_2xy$ pairs $(m,k)$ with $m\le x$ and $k\le y$.
\end{thm}
\begin{rmk}
It is not possible for such a lower bound to hold for all groups $G=G_{m, k}$.
As was noted in~\cite{BPS:2012}, several groups of this form do not arise in this way at all.
For example, the group $G_{11,1}$ never occurs as the group of points on any elliptic curve over any finite field.
\end{rmk}
Our final result for $M(G_{m, k})$ is that on average the full asymptotic of Theorem~\ref{meg-rephrased} holds unconditionally.
\begin{thm} \label{CDKS3} Fix $\epsilon>0$ and $A\ge1$. For $2\le x\le y^{1/4-\epsilon}$ we have that
\als{
\frac{1}{xy}\sum_{\substack{m\le x,\, k\le y \\ mk>1}} \left| M(G_{m,k}) - \frac{K(G_{m, k}) |G_{m, k}|^2}{|\Aut (G_{m, k})|\log|G_{m, k}|} \right|
\ll \frac{y}{(\log y)^A},
}
the implied constant depending at most on $A$ and $\epsilon$. Moreover, if the generalized Riemann hypothesis is true, then the same result is true for $x\le y^{1/2-\epsilon}$.
\end{thm}
In \cite{DS-MEN, DS-MEN-corr}, the second and fourth named authors studied the related question of how many elliptic curves over $\mathbb F_p$ have a given number of points, that is to say the asymptotic behaviour of
\[
M(N) := \sum_p \sum_{\substack{ E/\mathbb F_p\\ \#E(\mathbb F_p)=N}}\frac{1}{|\Aut_p(E)|} .
\]
It was shown in \cite{DS-MEN, DS-MEN-corr} that
\[
M(N)\sim K(N)\cdot \frac{N^2}{\phi(N)\log N} \quad(N\to\infty)
\]
under suitable assumptions on the distribution of primes in short arithmetic progressions, where
\eq{define K(N)}{
K(N) = \prod_{\ell\nmid N} \left(1- \frac{ \leg{N-1}{\ell}^2\ell+1}{(\ell-1)^2(\ell+1)} \right)
\prod_{\ell|N} \left(1-\frac{1}{\ell^{\nu_\ell(N)}(\ell-1)}\right).
}
Here $\nu_\ell(N)$ denotes the usual $\ell$-adic valuation of $N$.
As one might expect, the methods of this paper apply to the study of $M(N)$ as well.
We start by recording the obvious identity
\[
M(N) = \sum_{m^2k=N} M(G_{m, k}) .
\]
Then it is possible to show that, as expected, most of the contribution to $M(N)$ comes from groups $G_{m, k}$ with $m$ small, that is to say groups that are nearly cyclic.
\begin{thm}\label{MEN-1} For $N\ge1$ and $x\ge1$, we have that
\[
M(N) = \sum_{\substack{m^2k=N \\ m\le x}} M(G_{m, k})
+ O\left( \frac{N^2}{x\phi(N)\log(2N)} \right) .
\]
\end{thm}
Finally, we conclude with two more results on $M(N)$.
\begin{thm} \label{MEN-2} Let $N\ge1$ and set
\[
\eta = \frac{1}{N/(\log(2N))}
\sum_{\substack{ N^-<p\le N^+ \\ p\equiv 1\pmod m }} \sqrt{(p-N^-)(N^+-p)} ,
\]
and note that $\eta\ll1$ by the Brun-Titchmarsch inequality. For any fixed $\lambda>1$,
\[
\eta^\lambda \cdot \frac{N^2}{\phi(N) \log(2N) } \ll M(N) \ll \eta^{1/\lambda} \cdot \frac{N^2}{\phi(N)\log(2N)},
\]
the implied constants depending at most on $\lambda$.
\end{thm}
\begin{thm} \label{MEN-3} Fix $A>0$. For $x\ge1$, we have that
\als{
\frac{1}{x}\sum_{1<N\le x} \left| M(N) - \frac{K(N)N^2}{\phi(N)\log N} \right|
\ll_A \frac{x}{(\log x)^A} .
}
\end{thm}
The present paper also includes an appendix (by Greg Martin and the second and fourth named authors)
giving a probabilistic interpretation to the Euler factors arising in the constants $K(N)$ and $K(G)$ defined by~\eqref{define K(G)} and~\eqref{define K(N)}, respectively.
This interpretation is similar to the heuristic leading to the conjectural constants in related conjectures on properties of the reductions of a fixed global elliptic curve $E$ over the rationals
(e.g., the Lang-Trotter conjectures~\cite{LT:1976} and the Koblitz~\cite{Kob:1988} conjecture)
with the additional feature that the Euler factors at the primes $\ell$ dividing $N$ or $|G|$ are related to certain matrix counts over $\mathbb Z / \ell^e \mathbb Z$ for $e$ large enough.
\subsection*{Acknowledgements} The work of the second and third authors was partially supported by the Natural Sciences and Engineering Research Council of Canada.
Finally, part of this work was completed while the first, third and fourth authors were postdoctoral fellows at the Centre de recherches math\'ematiques at Montr\'eal, which they would like to thank for the financial support and the pleasant working environment.
\subsection*{Notation} Given a natural number $n$, we denote with $P^+(n)$ and $P^-(n)$ its largest and smallest prime factor, respectively, with the convention that $P^+(1)=1$ and $P^-(1)=\infty$. Moreover, we let $\tau_r(n)$ denote the coefficient of $1/n^s$ in the Dirichlet series $\zeta(s)^r$. In particular, $\tau_r(n)=r^{\omega(n)}$ for square-free integers $n$, where $\omega(n)$ denotes the number of distinct prime factors of $n$. In the special case when $r=2$, we simply write $\tau(n)$ in place of $\tau_2(n)$, which counts the number of divisors of $n$. We write $f*g$ to denote the Dirichlet convolution of the arithmetic functions $f$ and $g$, defined by $(f*g)(n)=\sum_{ab=n}f(a)g(b)$. As usual, given a Dirichlet character $\chi$, we write $L(s,\chi)$ for its Dirichlet series. In addition, we make use of the notation
\[
E(x,h;q) := \max_{(a,q)=1} \left| \sum_{\substack{ x <p \le x + h \\ p\equiv a\pmod q}} \log p - \frac{h}{\phi(q)} \right| .
\]
Finally, for $d\in\mathbb Z$ that is not a square and for $z\ge1$, we let
\[
\L(d) = L\left(1,\left(\frac{d}{\cdot}\right)\right) = \prod_{\ell} \left( 1- \frac{\leg{d}{\ell}}{\ell}\right)^{-1}
\quad\text{and}\quad
\L(d;z) = \prod_{\ell\le z} \left( 1- \frac{\leg{d}{\ell}}{\ell}\right)^{-1}.
\]
\section{Outline of the proofs}\label{outline}
In this section, we outline the chief ideas that go into the proofs of our main results.
However, most of our remarks concern the proofs of Theorems~\ref{CDKS} and~\ref{CDKS3}.
This is primarily because the remaining results are essentially corollaries of these theorems.
In particular, the main ingredient in the proof of Theorem~\ref{MEN-1} is Theorem~\ref{CDKS}, and the main ingredients in the proof Theorem~\ref{MEN-3} are Theorems~\ref{CDKS3} and~\ref{MEN-1} together with a short computation.
Theorem~\ref{MEN-2} is not truly a corollary, but its proof is essentially the same as that of Theorem~\ref{CDKS}.
The proof of Theorem~\ref{CDKS2} is somewhat different.
The ideas involved in its proof are essentially the same as those used to show Theorem 1.6 of~\cite{CDKS1} together with an application of Theorem~\ref{CDKS}.
All of this will be expounded further in Section~\ref{proofs}, where we complete the proofs of all six results.
For the remainder of this section, we focus our attention on outlining the main ingredients in the proofs of Theorems~\ref{CDKS} and~\ref{CDKS3}.
Throughout, we fix a group $G=G_{m,k}=\mathbb Z/m\mathbb Z\times\mathbb Z/mk\mathbb Z$, and we set $N=|G|=m^2k$. Moreover, given a prime $p\equiv 1\pmod {m}$, we set
\begin{eqnarray} \label{def-dp}
d_{m,k}(p)= \frac{(p-1-N)^2-4N}{m^2}=\left(\frac{p-1}{m}-mk\right)^2-4k.
\end{eqnarray}
Often, when the dependence on $m$ and $k$ is clear from the context, we will simply write $d(p)$ in place of $d_{k,m}(p)$.
Our starting point is the following lemma, whose proof is based on Deuring's work \cite{Deu} and its generalization due to Schoof \cite{Schoof}. We shall give the details of its proof in Section \ref{deuring lemma}.
\begin{lma}\label{formula for M(G)}
For any $m,k\in\mathbb N$, we have that
\[
M(G_{m, k}) = \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}\sum_{\substack{f^2\mid d(p) ,\, (f,k)=1 \\
d(p)/f^2\equiv 1,0\pmod 4 }}
\frac{\sqrt{|d(p)|} \L(d(p)/f^2)}{2\pi f} .
\]
\end{lma}
For the proof of Theorem \ref{CDKS}, we shall use the following simplified but weaker version of Lemma \ref{formula for M(G)}.
\begin{cor}\label{bounds for M(G)} For any $m,k\in\mathbb N$, we have that
\[
\sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}
\sqrt{|d(p)|} \L(d(p))
\ll M(G_{m, k}) \ll
\sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }} \frac{|d(p)|^{3/2}}{\phi(|d(p)|)}
\L(d(p)).
\]
\end{cor}
\begin{proof} For the lower bound, note that the term $f=1$ in Lemma \ref{formula for M(G)} always contributes to $M(G_{m, k})$, since $d(p)\equiv 0,1\mod 4$ for all $m,k$ and $p\equiv 1\mod m$. For the upper bound, notice that
\[
\L(d(p)/f^2) \le \frac{f}{\phi(f)} \L(d(p)) .
\]
Since
\[
\sum_{f|n} \frac{1}{\phi(f)}\ll \frac{n}{\phi(n)} ,
\]
the claimed upper bound follows.
\end{proof}
Evidently, Lemma \ref{formula for M(G)} and Corollary \ref{bounds for M(G)} reduce the estimation of $M(G_{m, k})$ to estimating an average of Dirichlet series evaluated at 1. In order to do so, we expand the Dirichlet series as an infinite sum and invert the order of summation by putting the sum over primes $p$ inside. For each fixed $n$ in the Dirichlet sum,
understanding this sum over primes involves understanding the distribution of the set
\eq{our set}{
\left\{\frac{p-1}{m}: N^-<p<N^+,\ p\equiv 1\mod m\right\}
}
in arithmetic progressions $a\mod b$, where the modulus $b=b(n)$ depends on $n$ and other parameters which are less essential.
Already when $b=m=1$, this problem is very hard and unsolved, even if we assume the validity of the Riemann Hypothesis. In order to limit the size of the moduli $b$ that are involved, we need to truncate the Dirichlet series that appear before inverting the order of summation. We could do this for each individual Dirichlet series, using character sum estimates such as the P\'olya-Vinogradov inequality or Burgess's bounds as in \cite{DS-MEN,DS-MEG}, but this would still leave us to deal with rather large moduli $b$. Instead, we use the following result,
which implies that for {\it most} characters $\chi$, $L(1,\chi)$ can be approximated by a very short Euler product, and then by a sum over integers $n$ supported only on small primes.
\begin{lma} \label{lemmashortproduct}
Let $\alpha \geq 1$ and $Q\ge3$. There is a set $\mathcal{E}_\alpha(Q)\subset[1,Q]\cap\mathbb Z$ of at most $Q^{2/\alpha}$ integers such that if $\chi$ is a Dirichlet character modulo $q\le\exp\{(\log Q)^2\}$ whose conductor does not belong to $\mathcal{E}_\alpha(Q)$, then
\[
L(1,\chi) = \prod_{\ell\le (\log Q)^{8\alpha^2}} \left(1-\frac{\chi(\ell)}{\ell}\right)^{-1} \left(1 + O_\alpha\left(\frac1{(\log Q)^\alpha}\right)\right).
\]
\end{lma}
\begin{proof} By a classical result, essentially due to Elliott (see \cite[Proposition 2.2]{GS}), we know that there is a set $\mathcal{E}_\alpha(Q)$ of at most $Q^{2/\alpha}$ integers from $[1,Q]$ such that
\[
L(1,\psi) = \prod_{\ell \leq (\log{Q})^{8\alpha^2}} \left( 1 - \frac{\psi(\ell)}{\ell} \right)^{-1}
\left( 1 + O\left(\frac{\alpha}{(\log Q)^\alpha}\right) \right)
\]
for all primitive characters $\psi$ of conductor in $[1,Q]\setminus\mathcal{E}_\alpha(Q)$. So if $\chi$ is a Dirichlet character modulo $q\le \exp\{(\log Q)^2\}$ induced by $\psi$ and the conductor of $\psi$ is in $[1,Q]\setminus \mathcal{E}_\alpha(Q)$, then
\als{
L(1, \chi)
&= \prod_{\ell \mid q} \left( 1 - \frac{\psi(\ell)}{\ell} \right)
\prod_{\ell \leq (\log{Q})^{8\alpha^2}} \left( 1 - \frac{\psi(\ell)}{\ell} \right)^{-1}
\left( 1 + O\left(\frac{\alpha}{(\log Q)^\alpha}\right) \right)\\
&= \prod_{\ell \mid q,\,\ell > (\log Q)^{8\alpha^2} } \left( 1 - \frac{\psi(\ell)}{\ell} \right)
\prod_{\ell \le (\log Q)^{8\alpha^2} } \left( 1 - \frac{\chi(\ell)}{\ell} \right)^{-1}
\left( 1 + O\left(\frac{\alpha}{(\log Q)^\alpha}\right) \right) .
}
Finally, note that
\[
\log\left(\prod_{\ell \mid q,\,\ell > (\log Q)^{8\alpha^2} }\left( 1 - \frac{\psi(\ell)}{\ell} \right) \right)
\ll \sum_{\ell \mid q,\,\ell > (\log Q)^{8\alpha^2} } \frac{1}{\ell}
\le \frac{\omega(q) }{(\log Q)^{8\alpha^2}} \ll \frac{1}{(\log Q)^{8\alpha^2-2}},
\]
since $\omega(q)\le \log q/\log 2\ll (\log Q)^2$, which completes the proof of the lemma.
\end{proof}
Expanding the short product in the above lemma leads to an approximation of $L(1,\chi)$ by a sum over $(\log Q)^A$-smooth integers, and we know that very few of them get $>Q^\epsilon$:
\begin{lma}\label{smooth} Let $f:\mathbb{N}\to\{z\in\mathbb C:|z|\le1\}$ be a completely multiplicative function. For $u\ge1$ and $x\ge10$ we have that
\[
\prod_{p\le x}\left(1-\frac{f(p)}p\right)^{-1}
= \sum_{\substack{P^+(n)\le x\\n\le x^u}} \frac{f(n)}n+ O\left(\frac{\log x}{e^u}\right) .
\]
\end{lma}
\begin{proof} We have that
\als{
\left| \prod_{p\le x}\left(1-\frac{f(p)}p\right)^{-1}
- \sum_{\substack{P^+(n)\le x\\n\le x^u}} \frac{f(n)}n \right|
= \left| \sum_{\substack{P^+(n)\le x\\ n > x^u }} \frac{f(n)}n \right|
&\le \frac{1}{e^u} \sum_{P^+(n)\le x} \frac{1}{n^{1-1/\log x}} \\
&\ll \frac{1}{e^u} \exp\left\{\sum_{p\le x} \frac{1}{p^{1-1/\log x}} \right\} .
}
So using the formula $p^{1/\log x}=1+O(\log p/\log x)$ and the prime number theorem, we obtain the claimed result.
\end{proof}
Combining Lemmas \ref{lemmashortproduct} and \ref{smooth}, we may replace $L(1,\chi)$ by a very short sum for most characters $\chi$, which means that we only need information for the distribution of the set \eqref{our set} for very small moduli. This leads to the following fundamental result, which is an improvement of Theorem \ref{meg-rephrased}. It will be proven in Section \ref{approx}.
\begin{thm}\label{approximation of M(G)} Fix $\alpha\ge1$ and $\epsilon\le 1/3$, and consider integers $m$ and $k$ with $1\le m\le k^{\alpha}$ and $k$ large enough so that
$k^{\frac{1}{2}-\epsilon} \ge(\log k)^{\alpha+2}$. Set $G=G_{m,k}$, and consider $h\in[mk^{\epsilon},m \sqrt{k}/(\log k)^{\alpha+2}]$. Then
\als{
M(G) &= \frac{K(G) |G|^2}{|\Aut (G)|\log|G|}
+ O_{\alpha,\epsilon} \left(\frac{k}{(\log k)^{\alpha}}
+ \frac{\sqrt{k}}{h} \sum_{q\le k^{\epsilon} } \tau_3(q)
\int_{N^-}^{N^+} E(y,h; qm ) \mathrm{d} y\right),
}
where $K(G)$ is defined by \eqref{define K(G)}.
\end{thm}
Even though we cannot estimate the error term for any given values of $m$ and $k$, we can do so if we average over $m$ and $k$ using the following result, which is a consequence of Theorem 1.1 in \cite{Kou}.
\begin{lma}\label{bv-short}Fix $\epsilon>0$ and $A\ge1$. For $x\ge h\ge2$ and $1\le Q^2\le h/x^{1/6+\epsilon}$, we have that
\[
\int_x^{2x}\sum_{q\le Q} E(y,h;q) \mathrm{d} y \ll\frac{xh}{(\log x)^A}.
\]
If, in addition, the Riemann hypothesis for Dirichlet $L$-functions is true, then the above estimate holds when $1\le Q^2\le h/x^\epsilon$.
\end{lma}
Theorem \ref{approximation of M(G)} and Lemma \ref{bv-short} lead to a proof of Theorem \ref{CDKS3} in a fairly straightforward way as we will see in Section \ref{proofs}.
Next, we turn to the proof of Theorem \ref{CDKS}.
Using Corollary~\ref{bounds for M(G)} and H\"older's inequality, we reduce the proof of this result to that of controlling sums of the form
\eq{CDKS - key sum}{
\sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}
\left(\frac{|d(p)|}{\phi(|d(p)|) }\right)^s \L(d(p))^r ,
}
where we take $r>0$ to prove the implicit upper bound and $r<0$ for the lower bound.
Nevertheless, we only seek an upper bound for the sum in \eqref{CDKS - key sum}, even for the lower bound in Theorem \ref{CDKS}. Therefore we can replace the sum over primes with a sum over almost primes and use sieve methods to detect the latter kind of integers. More precisely, we will majorize the characteristic function of primes $\le 2N$ by a convolution $\lambda*1$, where $\lambda$ is a certain truncation of the M\"obius function. This will be done using the {\it fundamental lemma of sieve methods}, which we state below in the form found in \cite[Lemma 5]{FI}. We could have also used Selberg's sieve, but the calculations are actually simpler when using Lemma \ref{lemma-FI}.
\begin{lma}\label{lemma-FI}
Let $y\ge2$ and $D=y^u$ with $u\ge2$. There exist two arithmetic functions $\lambda^\pm:\mathbb{N}\to[-1,1]$, supported on $\{d\in\mathbb{N}:P^+(d)\le y,\,d\le D\}$, for which
\[
\begin{cases}
(\lambda^-*1)(n)=(\lambda^+*1)(n)=1 &\text{if}\ P^-(n)>y,\\
(\lambda^-*1)(n)\le0\le(\lambda^+*1)(n) &\text{otherwise}.
\end{cases}
\]
Moreover, if $g:\mathbb{N}\to\mathbb{R}$ is a multiplicative function with $0\le g(p)\le\min\{2,p-1\}$ for all primes $p\le y$, and $\lambda\in\{\lambda^+,\lambda^-\}$, then
\[
\sum_d \frac{ \lambda(d)g(d) }{d}
= (1+O(e^{-u})) \prod_{p\le y} \left(1-\frac{g(p)}p\right).
\]
\end{lma}
Combining Lemmas \ref{lemmashortproduct} and \ref{lemma-FI}, we are led to following key result, which will proven in Section \ref{proof of Prop startwiththat}. As we will see in the same section, Theorem \ref{CDKS} is an easy consequence of this intermediate result.
\begin{prop}\label{startwiththat} Let $m,k\in\mathbb N$ and set $N=m^2k$. For any $r\in\mathbb{R}$ and $s\ge0$, we have that
\[
\sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}
\left(\frac{|d(p)|}{\phi(|d(p)|) }\right)^s \L(d(p))^r
\ll_{r,s} \left(\frac{k}{\phi (k)}\right)^r \frac{\sqrt{N}}{\phi(m)\log(2k)}.
\]
\end{prop}
\section{Completion of the proof of the main results}\label{proofs}
In this section we prove Theorems \ref{CDKS}-\ref{MEN-3}. We start by stating a preliminary result, which is Lemma 15 of \cite{DS-MEG} in slightly altered form.
\begin{lma}\label{Aut(G)} For $m,k\in\mathbb N$, we have that
\[
\frac{|{\Aut}(G_{m,k})|}{|G_{m,k}|}
= m\phi(m) \frac{\phi(k)}{k} \prod_{\substack{\ell | m \\ \ell\nmid k}} \left(1-\frac{1}{\ell^2} \right) .
\]
\end{lma}
\begin{proof}[Proof of Theorem \ref{CDKS}] The claimed inequalities are a consequence of Corollary \ref{bounds for M(G)}, Proposition \ref{startwiththat}, and H\"older's inequality.
Indeed, let $\mu=\lambda/(\lambda-1)$, so that $1/\lambda+1/\mu=1$. Then we have that
\als{
M(G_{m, k})
&\ll \sum_{\substack{ N^-<p<N^+ \\ p\equiv 1\mod m}} \sqrt{|d(p)|} \frac{|d(p)|}{\phi(|d(p)|)} \L(d(p)) \\
&\le \left( \sum_{\substack{ N^-<p<N^+ \\ p\equiv 1\mod m}} \sqrt{|d(p)|}\right)^{\frac{1}{\lambda}}
\left( \sum_{\substack{ N^-<p<N^+ \\ p\equiv 1\mod m}}
\sqrt{|d(p)|} \left(\frac{|d(p)|}{\phi(|d(p)|)}\right)^\mu \L(d(p))^{\mu}\right)^{\frac{1}{\mu}} \\
&\ll
\left( \sum_{\substack{ N^-<p<N^+ \\ p\equiv 1\mod m}} \frac{\sqrt{(N^+-p)(p-N^-)}}{m} \right)^{\frac{1}{\lambda}}
\left( \sum_{\substack{ N^-<p<N^+ \\ p\equiv 1\mod m}}
\sqrt{k} \left(\frac{|d(p)|}{\phi(|d(p)|)}\right)^\mu \L(d(p))^{\mu}\right)^{\frac{1}{\mu}} ,
}
since $|d(p)|=(N^+-p)(p-N^-)/m^2\ll N/m^2=k$. So the definition of $\delta$ and Proposition \ref{startwiththat} imply that
\[
M(G_{m, k}) \ll_{\lambda,A} \delta^{1/\lambda} \frac{km}{\phi(m)\log(2N)} \frac{k}{\phi(k)}.
\]
Hence the upper bound in Theorem \ref{CDKS} follows by Lemma \ref{Aut(G)}.
The proof of the lower bound is similar, having as a starting point the inequality
\[
\sum_{\substack{ N^-<p<N^+ \\ p\equiv 1\mod m}} \sqrt{|d(p)|}
\le \left( \sum_{\substack{ N^- <p \le N^+ \\ p\equiv 1\pmod{m} }}
\sqrt{|d(p)|} \L(d(p)) \right)^{\frac{1}{\lambda}}
\left( \sum_{\substack{ N^- <p \le N^+ \\ p\equiv 1\pmod{m} }}
\frac{\sqrt{|d(p)|}}{\L(d(p))^{\mu/\lambda}} \right)^{\frac{1}{\mu}} .
\]
\end{proof}
\begin{proof}[Proof of Theorem \ref{MEN-2}]
The proof of Theorem \ref{MEN-2} is completely analogous to the proof of Theorem \ref{CDKS}. The only difference is that instead of starting with Corollary \ref{bounds for M(G)}, we observe that
\[
\sum_{N^-<p<N^+} \sqrt{|D_N(p)|} \L(D_N(p))
\ll M(N) \ll
\sum_{N^-<p<N^+} \frac{|D_N(p)|^{3/2}}{\phi(|D_N(p)|)}
\L(D_N(p)),
\]
a consequence of relation~\eqref{reduction to class number avg} below with $n=1$.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{CDKS2}] Note that when $m=k=1$ and $N=1$, then $N^+=4$ and $N^-=0$ and thus the primes 2 and 3 belong to the set $\{N^-<p\le N^+: p\equiv 1\pmod m\}$. So, by Theorem \ref{CDKS}, it suffices to show Theorem \ref{CDKS2} when $y$ is large enough. We further assume that $x\in\mathbb N$, which we may certainly do. Observe that $(N^+-p)(p-N^-)\asymp N$ for $p\in((\sqrt{N}-1/2)^2,(\sqrt{N}+1/2)^2)$, and thus
\[
\frac{1}{N/(\phi(m)\log(2N))} \sum_{\substack{N^-<p<N^+ \\ p\equiv 1\mod{p} }} \sqrt{ (N^+-p)(p-N^-)}
\gg \frac{\phi(m)}{\sqrt{N}} \sum_{\substack{(\sqrt{N}-1/2)^2<p<(\sqrt{N}+1/2)^2 \\ p\equiv 1\mod{m} }} \log p .
\]
So, if we set
\[
C(m,k) = \frac{|G_{m, k}|^2}{|\Aut(G_{m, k})| \log(2G_{m, k})} \asymp \frac{mk^2}{\phi(m)\phi(k)\log(mk)},
\]
then Theorem \ref{CDKS} with $\lambda=2$ implies that
\als{
\sum_{\substack{3x/4<m\le x \\ y/100<k\le y}} \sqrt{ \frac{M(G_{m, k})}{C(m,k)} }
&\gg \sum_{\substack{3x/4<m\le x \\ y/100<k\le y}} \frac{\phi(m)}{x\sqrt{y}}
\sum_{\substack{ (m\sqrt{k}-1/2)^2<p< (m\sqrt{k}+1/2)^2 \\ p\equiv 1\pmod m}} \log p \\
&\ge\sum_{3x/4<m\le x} \sum_{\substack{ x^2y/3 < p \le 4x^2y/9 \\ p\equiv 1\pmod m }} \frac{\phi(m)\log p}{x\sqrt{y}}\sum_{ \substack{y/100<k \le y \\ (\sqrt p -1/2)^2/m^2<k < (\sqrt p+1/2)^2/m^2 }} 1 ,
}
provided that $y$ is large enough. Note that
\[
\frac{(\sqrt p+1/2)^2 - (\sqrt p-1/2)^2}{m^2} = \frac{2\sqrt{p}}{m^2} \ge \frac{2x\sqrt{y/3}}{x^2} > 1,
\]
by our assumptions that $x\le\sqrt{y}$. Since we also have that $(\sqrt p-1/2)^2/m^2>y/100$ and that $(\sqrt p+1/2)^2/m^2\le y$ for $y$ large enough and $m$ and $p$ as above, we conclude that
\[
\sum_{\substack{3x/4<m\le x \\ y/100<k\le y}} \sqrt{ \frac{M(G_{m, k})}{C(m,k)} }
\gg \frac{1}{x^2} \sum_{3x/4<m\le x} \phi(m)
\sum_{\substack{ x^2y/3 < p \le 4x^2y/9 \\ p\equiv 1\pmod m }} \log p.
\]
This last double sum equals
\[
\sum_{3x/4<m\le x} \phi(m) \cdot
\frac{x^2y}{9\phi(m)} + O_A\left( \frac{x^3y}{(\log y)^A}\right) \gg x^3y,
\]
by the Bombieri Vinogradov theorem. Therefore we conclude that
\[
\sum_{\substack{3x/4<m\le x \\ y/100<k\le y}} \sqrt{ \frac{M(G_{m,k})}{C(m,k)} } \gg xy.
\]
Since the summands are all $\ll1$ in this range by Theorem \ref{CDKS} (recall that $\delta\ll1$ there), we obtain Theorem \ref{CDKS2}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{CDKS3}] Let $\theta$ be a parameter, which we take to be $1/2$ or $1/4$, according to whether we assume the generalized Riemann hypothesis or not. We then suppose that $1\le x\le y^{\theta-\epsilon}$. Note that Theorem \ref{CDKS} and Lemma \ref{Aut(G)} imply that
\[
\sum_{\substack{m\le x,\, k\le y/(\log y)^A \\ mk>1}} \left| M(G_{m,k}) - \frac{K(G_{m, k}) |G_{m, k}|^2}{|\Aut (G_{m, k})|\log|G_{m, k}|} \right|
\ll \frac{xy^2}{(\log y)^A} .
\]
We break the remaining range of $m$ and $k$ into dyadic intervals, hence reducing Theorem \ref{CDKS3} to showing that
\[
E:= \sum_{\substack{ x/2< m\le x \\ y/2< k\le y}}
\left| M(G_{m,k}) - \frac{K(G_{m, k}) |G_{m, k}|^2}{|\Aut (G_{m, k})|\log|G_{m, k}|} \right| \ll_{\epsilon,A} \frac{xy^2}{(\log y)^A}
\]
for $x\le y^{\theta-\epsilon}$. (Note that these might be different values of $x,y$ and $\epsilon$ than the ones we started with.) We apply Theorem \ref{approximation of M(G)} with $h= (x^2y)^{1/2}/(\log y)^{A+2}$ for all $m\in[x/2,x]$ and $k\in[y/2,y]$, to deduce that
\als{
E &\ll \frac{\sqrt{y}}{h} \sum_{\substack{ x/2< m\le x \\ y/2< k\le y}} \sum_{q \le k^{\epsilon} } \tau_3(q)
\int_{(m^2k)^-}^{(m^2k)^+} E(t,h; qm ) \mathrm{d} t
+ \frac{xy^2}{(\log y)^A} \\
&=: E'+ \frac{xy^2}{(\log y)^A} ,
}
say. Putting the sum over $k$ inside, we find that
\als{
E'&\ll \frac{\sqrt{y}}{h} \sum_{x/2<m\le x} \sum_{q \le y^{\epsilon} } \tau_3(q)
\int_{x^2y/10}^{2x^2y} E(t,h; qm)
\left(\sum_{ \substack{ y/2<k\le y \\ t^-/m^2<k < t^+/m^2}} 1 \right) \mathrm{d} t \\
&\ll \frac{y}{hx} \sum_{m\le x} \sum_{q \le y^{\epsilon} } \tau_3(q)
\int_{x^2y/10}^{2x^2y} E(t,h; qm) \mathrm{d} t
\le \frac{y}{hx} \sum_{m\le x} \sum_{q \le y^{\epsilon} } \tau_4(q)
\int_{x^2y/10}^{2x^2y} E(t,h; q) \mathrm{d} t .
}
We note that $E(u,h; b) \ll \sqrt{h/\phi(b)} \sqrt{E(u,h;b)}$, by the Brun-Titchmarsch inequality. So the Cauchy-Schwarz inequality and Lemma \ref{bv-short} imply that
\als{
E'&\ll \frac{y}{x h} \left(\sum_{b\le xy^{3\epsilon}} \tau_4(b)^2 \int_{x^2y/10}^{2x^2y}
\frac{h}{\phi(b)} \mathrm{d} t\right)^{\frac{1}{2}}
\left( \sum_{b\le xy^{3\epsilon}} \int_{x^2y/10}^{2x^2y} E(t,h; b) \mathrm{d} t\right)^{\frac{1}{2}} \\
&\ll \frac{y}{x h} \left( x^2yh(\log y)^{16} \cdot \frac{x^2yh}{(\log y)^{2A+16}} \right)^{\frac{1}{2}}
= \frac{xy^2}{(\log y)^A} ,
}
which completes the proof of Theorem \ref{CDKS3}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{MEN-1}]
Theorem~\ref{CDKS} implies that
\[
M(G_{m, k}) \ll \frac{k^{3/2}}{\phi (k)} \frac{\sqrt{N}}{\phi(m)\log(2k)} = \frac{mk^2}{\phi(k)\phi(m)\log(2k)}\le \frac{Nmk}{\phi(N)\phi(m)\log(2k)}.
\]
Therefore,
\[
\sum_{\substack{m^2k=N \\ m>x}} M(G_{m, k})
\ll \sum_{\substack{m^2|N \\ x<m\le\sqrt{N} }} \frac{N^2}{m\phi(m)\phi(N)\log(2N/m^2)}
\ll \frac{N^2}{x\phi(N)\log(2N)},
\]
which completes the proof of Theorem \ref{MEN-1}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{MEN-3}]
In view of Theorem \ref{MEN-1}, it suffices to show that
\[
\sum_{1<N\le x} \left| \sum_{\substack{ m^2k=N \\ m\le (\log x)^A}} M(G_{m, k}) - \frac{K(N)N^2}{\phi(N)\log N} \right|
\ll_A \frac{x^2}{(\log x)^A},
\]
where $K(N)$ is defined by~\eqref{define K(N)}.
Note that
\als{
&\sum_{1<N\le x} \left| \sum_{\substack{ m^2k=N \\ m\le (\log x)^A}} M(G_{m, k})
- \sum_{\substack{ m^2k=N \\ m\le (\log x)^A}} \frac{K(G_{m, k}) |G_{m, k}|^2}{|\Aut (G_{m, k})|\log|G_{m, k}|} \right| \\
&\quad\le \sum_{\substack{1<m^2k\le x \\ m\le (\log x)^A}}
\left| M(G_{m, k}) - \frac{K(G_{m, k}) |G_{m, k}|^2}{|\Aut (G_{m, k})|\log|G_{m, k}|} \right| \\
&\quad\le \sum_{1\le 2^j\le (\log x)^{A} }
\sum_{\substack{k\le x/4^j \\ 2^j\le m<2^{j+1} \\ m^2k>1 }}
\left| M(G_{m, k}) - \frac{K(G_{m, k}) |G_{m, k}|^2}{|\Aut (G_{m, k})|\log|G_{m, k}|} \right| \\
&\quad\ll_A \sum_{1\le 2^j\le (\log x)^{A}} \frac{x^2}{8^j(\log x)^A} \ll \frac{x^2}{(\log x)^A}
}
by Theorem~\ref{CDKS3}.
So it suffices to show that
\eq{MEN-3 goal}{
\sum_{1<N\le x} \frac{N}{\log N} \left| \sum_{\substack{ m^2k=N \\ m\le (\log x)^A}} \frac{K(G_{m, k}) |G_{m, k}|}{|\Aut (G_{m, k})|}
- \frac{K(N)N}{\phi(N)} \right|
\ll_A \frac{x^2}{(\log x)^A} .
}
In fact, Lemma \ref{Aut(G)} implies that
\als{
\frac{ K(G_{m, k}) |G_{m, k}|}{|\Aut(G_{m, k})|}
&= \frac{k}{m\phi(m)\phi(k)} \prod_{\substack{\ell | m \\ \ell\nmid k}} \left(1-\frac{1}{\ell^2} \right)^{-1} K(G_{m, k}) \\
&= \frac{N}{m^2\phi(N)}\prod_{\ell|(m,k)} \left(1-\frac{1}{\ell}\right)^{-1}
\prod_{\substack{\ell | m \\ \ell\nmid k}} \left(1-\frac{1}{\ell^2} \right)^{-1} K(G_{m, k}) \\
&= \frac{N}{m^2\phi(N)} \prod_{\ell\nmid N}\left(1-\frac{\leg{N-1}{\ell}^2\ell+1}{(\ell-1)^2(\ell+1)}\right)
\prod_{\ell\mid (m,k)}\left(1+\frac{1}{\ell}\right)
\prod_{\substack{\ell\mid k\\ \ell\nmid m}}\left(1-\frac{1}{\ell(\ell-1)}\right) .
}
Therefore,
\als{
\sum_{\substack{m^2k = N \\ m\le (\log x)^A}} \frac{K(G_{m, k}) |G_{m, k}|}{|\Aut(G_{m, k})|}
&= \sum_{m^2k = N} K(G_{m, k}) \frac{|G_{m, k}|}{|\Aut(G_{m, k})|} + O\left(\frac{N}{(\log x)^A\phi(N)}\right) \\
&= \frac{N}{\phi(N)} \prod_{\ell\nmid N}\left(1-\frac{\leg{N-1}{\ell}^2\ell+1}{(\ell-1)^2(\ell+1)}\right) \cdot S(N)
+ O\left(\frac{N}{(\log x)^A\phi(N)}\right),
}
where
\[
S(N) = \sum_{m^2k=N} \frac{1}{m^2} \prod_{\ell\mid (m,k)}\left(1+\frac{1}{\ell}\right)
\prod_{\substack{\ell\mid k\\ \ell\nmid m}}\left(1-\frac{1}{\ell(\ell-1)}\right) .
\]
Note that
\als{
S(\ell^v) &= 1-\frac{1}{\ell(\ell-1)} + \sum_{1\le j\le v/2} \frac{1}{\ell^{2j}} \left(1+\frac{{\bf 1}_{j<v/2}}{\ell}\right) \\
&=1-\frac{1}{\ell(\ell-1)} + \sum_{1\le j\le v/2} \frac{1}{\ell^{2j}}
+ \sum_{1\le j\le v/2} \frac{{\bf 1}_{j<v/2}}{\ell^{2j+1}} \\
&=1-\frac{1}{\ell(\ell-1)} + \sum_{i=2}^v \frac{1}{\ell^i} = 1-\frac{1}{\ell^v(\ell-1)} .
}
So we conclude that
\[
\sum_{\substack{m^2k = N \\ m\le (\log x)^A}} \frac{K(G_{m, k}) |G_{m, k}|}{|\Aut(G_{m, k})|}
= \frac{K(N) N}{\phi(N)} + O\left(\frac{N}{(\log x)^A\phi(N)}\right) ,
\]
which yields relation \eqref{MEN-3 goal}, thus completing the proof of Theorem \ref{MEN-3}.
\end{proof}
\section{Reduction to an average of Dirichlet series}\label{deuring lemma}
In this section, we prove Lemma \ref{formula for M(G)} using the theory developed by Deuring~\cite{Deu} and somewhat generalized by Schoof~\cite{Schoof}. As before, we fix a group $G=G_{m,k}=\mathbb Z/m\mathbb Z\times\mathbb Z/mk\mathbb Z$, and we set $N=|G|=m^2k$. Given a prime $p$ and an integer $n$ such that $n^2|N$, we define
\[
M_p(N;n)=\sum_{\substack{E/\mathbb F_p\\ \#E(\mathbb F_p)=N\\ E(\mathbb F_p)[n]\cong G_{n,1}}}\frac{1}{|\Aut_p(E)|},
\]
the weighted number of isomorphism classes of elliptic curves over any prime finite field which have exactly $N$ rational points and whose rational $n$-torsion subgroup is isomorphic to $G_{n,1}=\mathbb Z/n\mathbb Z\times\mathbb Z/n\mathbb Z$.
It is not hard to relate $M_p(G)$ to a sum involving $M_p(N;n)$.
This is accomplished via an inclusion-exclusion argument, which gives the relation
\begin{equation}\label{inclusion exclusion}
M_p(G)=\sum_{r^2 | k} \mu(r)M_p(N; r m).
\end{equation}
In~\cite{Schoof}, Schoof essentially gave a formula for $M_p(N;n)$ in terms of class numbers.
However, one needs to exercise care here as Schoof counts each $\mathbb F_p$-isomorphism class $E$ with weight $1$ instead of with weight $1/|\Aut_p(E)|$ as we do here.
Given a negative discriminant $D$, we let $H(D)$ denote the \textit{Kronecker class number}, which is defined as
\[
H(D)=\sum_{ \substack{f^2\mid D\\ D/f^2 \equiv 0,1\pmod{4} }} \frac{h(D/f^2)}{w(D/f^2)}.
\]
Here, as usual, $h(d)$ denotes the (ordinary) class number of the unique imaginary quadratic order of discriminant $d$,
and $w(d)$ denotes the cardinality of its unit group.
Then letting
\[
D_N(p)=(p+1-N)^2-4p=(p-1-N)^2-4N
\]
and reworking the proofs of~\cite[Lemma 4.8 and Theorem 4.9]{Schoof} to count each class $E$ with weight $1/|\Aut_p(E)|$,
we arrive at the formula
\eq{reduction to class number avg}{
M_p(N;n)=
\begin{cases}
H\left(\frac{D_N(p)}{n^2}\right)&\text{if }p\in(N^-,N^+)\text{ and }p\equiv 1\pmod n,\\
0&\text{otherwise}.
\end{cases}
}
Note here that $D_N(p)/n^2$ is a negative discriminant whenever $p\in(N^-,N^+)$, $p\equiv 1\pmod n$, and $n^2\mid N$.
\begin{lma}\label{Deuring for groups}
Let $m,k\in\mathbb N$
and recall that $d(p) = d_{m,k}(p)$ is defined by \eqref{def-dp}. If $p\in (N^-,N^+)$ and $p\equiv 1\pmod m$, then
\[
M_p(G_{m, k})=
\sum_{\substack{f^2\mid d(p),\, (f,k)=1\\ \frac{d(p)}{f^2}\equiv 0,1\pmod 4}}\frac{h(d(p)/f^2)}{w(d(p)/f^2)}.
\]
Otherwise, $M_p(G_{m, k})=0$.
\end{lma}
\begin{rmk}
The above formula is amenable to computation.
Indeed, given a prime $p$ and any $m$ and $k$,
very simple modifications to the usual quadratic forms algorithm for computing class numbers (see~\cite[pp.~99--100]{BV:2007} for example)
make it possible to compute $M_p(G_{m, k})$ using at most $O(k)$ arithmetic operations, which is reasonable for small $k$.
If we put
\begin{equation*}
H_k(D)=\sum_{\substack{f^2\mid D,\, (f,k)=1\\ \frac{D}{f^2}\equiv 0,1\pmod 4}}\frac{h(D/f^2)}{w(D/f^2)}
\end{equation*}
for each negative discriminant $D$ and each positive integer $k$, then the only modifications needed are as follows.
When the algorithm produces the (not necessarily primitive) form $ax^2+bxy+cy^2$, say with $(a,b,c)=f\ge1$,
it is counted subject to the following rules, provided that $(f,k)=1$.
\begin{enumerate}
\item Forms proportional to $x^2+y^2$ are counted with weight $1/4$.
\item Forms proportional to $x^2+xy+y^2$ are counted with weight $1/6$.
\item All other forms are counted with weight $1/2$.
\end{enumerate}
Similarly, tables of $M(G_{m, k})$ or $M_p(G_{m, k})$ values can be computed for $m$ and $k$ of modest size by simultaneously computing a table of values of $H_k(D)$.
\end{rmk}
\begin{proof}
It follows from~\eqref{reduction to class number avg} that $M_p(G)=0$ unless $p\in(N^-,N^+)$ and $p\equiv 1\pmod m$.
Therefore, assume that $p\in (N^-,N^+)$ and $p\equiv 1\pmod m$, and write $k=s^2t$ with $t$ square-free.
Combining relations~\eqref{inclusion exclusion} and~\eqref{reduction to class number avg}
with the definition of the Kronecker class number, we find that
\begin{equation*}
\begin{split}
M_p(G)
=\sum_{\substack{r\mid s\\ p\equiv 1\pmod{rm}}}\mu(r)H\left(\frac{D_N(p)}{(rm)^2}\right)
&=\sum_{\substack{r\mid s\\ p\equiv 1\pmod{rm}}}\mu(r)H\left(\frac{d(p)}{r^2}\right)\\
&=\sum_{\substack{r\mid s\\ p\equiv 1\pmod{rm}}}\mu(r)
\sum_{\substack{f^2\mid\frac{d(p)}{r^2}\\ \frac{d(p)}{(rf)^2}\equiv 0,1\pmod 4}}
\frac{h(d(p)/(rf)^2)}{w(d(p)/(rf)^2)}\\
&=\sum_{\substack{r\mid s\\ p\equiv 1\pmod{rm}}}\mu(r)
\sum_{\substack{f^2\mid d(p),\, r\mid f\\ \frac{d(p)}{f^2}\equiv 0,1\pmod 4}}\frac{h(d(p)/f^2)}{w(d(p)/f^2)}.
\end{split}
\end{equation*}
Now interchanging the sum over $r$ with the sum over $f$ and recalling the identity
\begin{equation*}
\sum_{r\mid n}\mu(n)=\begin{cases}1&\text{if }n=1,\\ 0&\text{otherwise},\end{cases}
\end{equation*}
we arrive at the formula
\begin{equation*}
\begin{split}
M_p(G)
&=
\sum_{\substack{f^2\mid d(p)\\ (f,s,(p-1)/m)=1\\ \frac{d(p)}{f^2}\equiv 0,1\pmod 4}}\frac{h(d(p)/f^2)}{w(d(p)/f^2)}.
\end{split}
\end{equation*}
In order to complete the proof, it is sufficient to show that, in the above sum,
the condition $(f,s,(p-1)/m)=1$ implies the simpler condition $(f,k)=1$, the converse implication being immediate.
To this end, we write $p=1+jm$ and assume that $(f,s,(p-1)/m)=(f,s,j)=1$.
Then $d(p)=(j-mk)^2-4k$, and the condition $d(p)/f^2\equiv 0, 1\pmod 4$ may be rewritten as
\begin{equation}\label{disc condition}
(j-mk)^2-4k\equiv 0, f^2\pmod{4f^2}.
\end{equation}
Now let $\ell$ be any prime dividing $(f,k)$.
Then the above congruence implies that $\ell\mid j$, but that implies that $\ell^2\mid (j-mk)^2$.
Whence $\ell^2\mid 4k$.
If $\ell$ is odd, then we have that $\ell^2\mid k$, and hence $\ell\mid (f,s,j)=1$, which is a contradiction.
If $\ell=2$, then we divide~\eqref{disc condition} through by $4$ to obtain
\begin{equation*}
\left(\frac{j}{2}-m\frac{k}{2}\right)^2-k\equiv 0, \frac{f^2}{4}\pmod{f^2}.
\end{equation*}
Since $\ell=2\mid (f,k)$,
we have that $k$ is even and congruent to a difference of two squares modulo $4$.
This in turn implies that $k\equiv 0\pmod 4$, i.e., $2\mid s$.
Thus, in this case we also have the contradiction $\ell=2\mid (f,s,j)=1$.
Therefore, we conclude that $(f,k)=1$, and this completes the proof of the lemma.
\end{proof}
Lemma~\ref{Deuring for groups} together with the class number formula immediately yields Lemma \ref{formula for M(G)}.
\section{Local computations}\label{local computations}
In this section we gather some local computations which we will need in the proofs of Theorem \ref{approximation of M(G)} and Proposition \ref{startwiththat}. As before, we continue to assume that $m, k,$ and $N$ are positive integers with $N=|G_{m, k}|=m^2k$.
\begin{lma}\label{generic quad lemma}
Let $\ell$ be an odd prime prime. For $e\ge1$, $(d,\ell)=1$ and $(a,b)=1$, we have that
\[
\#\{j\in\mathbb Z/\ell^e\mathbb Z : j^2\equiv d\pmod{\ell^e}\} = 1+ \leg{d}{\ell}
\]
and
\[
\#\{j\in\mathbb Z/\ell^e\mathbb Z : j^2\equiv d\pmod{\ell^e},\,(a+bj,\ell)=1\} = 1+ \leg{a^2-db^2}{\ell}^2 \leg{d}{\ell} .
\]
\end{lma}
\begin{proof} The first formula is classical. For the second, we first note that if $\leg{d}{\ell}=-1$, then $\leg{a^2-db^2}{\ell}^2=1$, and the formula holds. Now assume that $\leg{d}{\ell}=1$, so that there are exactly two solutions to the congruence $j^2\equiv d\mod{\ell^e}$, say $\pm j_0$. If $\ell\mid b$, then the condition $(a+bj,\ell)=1$ is satisfied trivially for all $j\in\mathbb Z$, and the claimed result follows. Finally, if $\ell\nmid b$, then we need to exclude exactly one of the solutions when $a\equiv \pm bj_0\mod{\ell}$, that is to say when $a^2\equiv b^2d\mod{\ell}$. So the claimed formula holds in this last case too.
\end{proof}
We set
\eq{T(n) def}{
T(n) = \sum_{d\mod n} \leg{d-4k}{n} \#\{j\mod n: j^2\equiv d\mod n,\ (N+1+jm,n)=1\}.
}
\begin{prop}\label{odd primes prop} Let $\ell$ be a prime not dividing $2k$ and $w\ge1$. Then
\[
\frac{T(\ell^w)}{\ell^{w-1}} = - \leg{m(N-1)}{\ell}^2 +
\begin{cases}
\ell - 1 - \leg{k}{\ell}
&\mbox{if $w$ is even},\cr
- 1
&\mbox{if $w$ is odd}.
\end{cases}
\]
\end{prop}
\begin{proof} We write $T(\ell^w)=T_1(\ell^w)+T_2(\ell^w)$, where $T_1(\ell^w)$ is the same sum as $T(\ell^w)$ with the additional restriction that $\ell|d$ and $T_2(\ell^w)$ is the remaining sum. First, we calculate $T_1(\ell^w)$. We have that
\als{
T_1(\ell^w)
&= \sum_{\substack{d\mod{\ell^w}\\\ell|d}}
\leg{d-4k}{\ell^w} \sum_{\substack{j\mod{\ell^w} \\ j^2\equiv d\mod {\ell^w}}}\leg{N+1+jm}{\ell}^2 \\
&= \sum_{\substack{d\mod{\ell^w}\\\ell|d}}
\leg{-4k}{\ell}^w \leg{N+1}{\ell}^2 \sum_{\substack{j\mod{\ell^w},\, \ell|j \\ j^2\equiv d\mod {\ell^w}}} 1 \\
&= \leg{-k}{\ell}^w \leg{N+1}{\ell}^2 \sum_{\substack{j\mod{\ell^w} \\ \ell|j }} 1
= \leg{-k}{\ell}^w \leg{N+1}{\ell}^2 \ell^{w-1} .
}
Finally, we compute $T_2(\ell^w)$. Applying Lemma \ref{generic quad lemma}, we find that
\als{
T_2(\ell^w) &= \sum_{\substack{d\mod{\ell^w}\\ (d,\ell)=1}} \leg{d-4k}{\ell}^w
\left( 1+\leg{(N+1)^2-dm^2}{\ell}^2 \leg{d}{\ell} \right) \\
&= \ell^{w-1} \sum_{d\mod{\ell}} \leg{d-4k}{\ell}^w
\left( 1+\leg{(N+1)^2-dm^2}{\ell}^2 \leg{d}{\ell} \right) - \ell^{w-1}\leg{-k}{\ell}^w.
}
If $\ell\mid m$, then $\leg{(N+1)^2-dm^2}{\ell}=1$ for all $d\mod\ell$. On the other hand, if $\ell \nmid m$, then there is precisely one $d\mod\ell$ such that $(N+1)^2-dm^2\equiv 0\mod{\ell}$, for which we have that
\[
\leg{d-4k}{\ell}^{w}=\leg{m^2d-4m^2k}{\ell}^{w} = \leg{(N-1)^2}{\ell}^{w} = \leg{N-1}{\ell}^2
\quad\text{and}\quad
\leg{d}{\ell} = \leg{N+1}{\ell}^2.
\]
Thus, whether $\ell$ divides $m$ or not, we have
\[
\frac{T_2(\ell^w)}{\ell^{w-1}} = - \leg{-k}{\ell}^w - \leg{m(N-1)(N+1)}{\ell}^2
+ \sum_{d\mod\ell} \leg{d-4k}{\ell}^{w} \left( 1+ \leg{d}{\ell} \right),
\]
which implies that
\als{
\frac{T(\ell^w)}{\ell^{w-1}} & = \leg{-k}{\ell}^w \leg{N+1}{\ell}^2 - \leg{-k}{\ell}^w - \leg{m(N-1)(N+1)}{\ell}^2 \\
&\quad + \sum_{d\mod\ell} \leg{d-4k}{\ell}^{w} \left( 1+ \leg{d}{\ell} \right) .
}
Note that if $\ell|N+1$, then $\leg{-k}{\ell}=1$ and thus
\[
\leg{-k}{\ell}^w \leg{N+1}{\ell}^2 - \leg{-k}{\ell}^w - \leg{m(N-1)(N+1)}{\ell}^2 = -1 = - \leg{m(N-1)}{\ell}^2,
\]
whereas if $\ell\nmid N+1$, then
\[
\leg{-k}{\ell}^w \leg{N+1}{\ell}^2 - \leg{-k}{\ell}^w - \leg{m(N-1)(N+1)}{\ell}^2 = - \leg{m(N-1)}{\ell}^2 .
\]
So
\als{
\frac{T(\ell^w)}{\ell^{w-1}}
&= - \leg{m(N-1)}{\ell}^2
+ \sum_{d\mod\ell} \leg{d-4k}{\ell}^{w} \left( 1+ \leg{d}{\ell} \right) .
}
If now $w$ is odd, then
\[
\sum_{d\mod\ell} \leg{d-4k}{\ell}^{w} \left( 1+ \leg{d}{\ell} \right)
= \sum_{d\mod\ell} \leg{d-4k}{\ell}\leg{d}{\ell}
= -1 ,
\]
using for example \cite[Exercise 1.1.9]{Stepanov} since $(2k, \ell)=1$.
Finally, if $w$ is even, then
\[
\sum_{d\mod\ell} \leg{d-4k}{\ell}^{w} \left( 1+ \leg{d}{\ell} \right)
= \ell-1 + \sum_{\substack{d\mod\ell \\ d\not\equiv 4k\mod{\ell}}} \leg{d}{\ell}
= \ell-1 - \leg{k}{\ell} ,
\]
which completes the proof of the proposition.
\end{proof}
\begin{cor}\label{formula for P(ell)} For a prime $\ell$ not dividing $2k$, we have that
\[
P(\ell) := 1 + \sum_{w\ge1} \frac{T(\ell^w)}{\ell^{2w-1}(\ell- \leg{m}{\ell}^2) }
= \frac{\ell^3-\leg{m}{\ell}^2\ell^2 - (1+\leg{m}{\ell}^2\leg{N-1}{\ell}^2)\ell - 1- \leg{N-1}{\ell}^2\leg{k}{\ell} }
{(\ell^2-1)(\ell-\leg{m}{\ell}^2)} .
\]
\end{cor}
\begin{proof}
Lemma \ref{odd primes prop} and a straightforward computation imply that
\[
P(\ell) = \frac{\ell^3-\leg{m}{\ell}^2\ell^2 - (1+\leg{m}{\ell}^2\leg{N-1}{\ell}^2)\ell + \leg{m}{\ell}^2-\leg{m(N-1)}{\ell}^2-1-\leg{k}{\ell} } {(\ell^2-1)(\ell-\leg{m}{\ell}^2)} .
\]
Finally, note that
\[
\leg{m(N-1)}{\ell}^2+\leg{k}{\ell} - \leg{m}{\ell}^2 = \leg{N-1}{\ell}^2\leg{k}{\ell} ,
\]
since $\leg{k}{\ell}=\leg{m}{\ell}^2=1$ if $\ell|N-1$.
\end{proof}
\section{Proof of Proposition \ref{startwiththat}}\label{proof of Prop startwiththat}
This section is dedicated to the proof of Proposition \ref{startwiththat}, which gives an upper bound of the conjectured order of magnitude for the average of special values
\[
\L(d(p)) = L \left(1, \displaystyle{\left( \frac{d(p)}{\cdot}\right)} \right)
\]
summed over integers with no small prime factors. A key role will be played by the fundamental lemma of sieve methods, i.e. Lemma \ref{lemma-FI}.
\begin{proof}[Proof of Proposition \ref{startwiththat}] We shall employ the notation
\[
\rho(n) := \frac{|n|}{\phi(|n|)} = \prod_{\ell | n} \left(1-\frac{1}{\ell}\right)^{-1} .
\]
We will simplify the sum we are estimating with an application of the Cauchy-Schwarz inequality but, first, we massage the
$L$-functions that appear in it. Note that if $p=1+jm$, then $d(p)=(j-mk)^2 - 4k \equiv j^2\pmod k$.
So
\als{
\L(d(p))^r
= \prod_{ \substack{ \ell | k \\ \ell\nmid j}} \left(1-\frac{1}{\ell} \right)^{-r}
\prod_{ \ell \nmid k} \left(1-\frac{ \leg{d(p)}{\ell} }{\ell}\right)^{-r}
\ll_r \rho(k)^r \rho((j,k))^{|r|} \L(k^2d(p))^r,
}
and consequently,
\[
S := \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}
\rho(d(p))^{s} \L(d(p))^r
\ll_r \rho(k)^r \sum_{\substack{N^-<p<N^+ \\ p =1+jm,\, j\in\mathbb N }} \rho((j,k))^{|r|} \rho(d(p))^s
\L(k^2d(p))^r.
\]
Hence the Cauchy-Schwarz inequality yields that
\eq{C-S}{
\frac{S}{\rho(k)^r }
\ll_r \left(\sum_{\substack{N^-<p<N^+ \\ p =1+jm }} \rho((j,k))^{2|r|} \rho(d(p))^{2s} \right)^{\frac{1}{2}}
\left(\sum_{\substack{N^-<p<N^+ \\ p\equiv 1\pmod m }}
\L(k^2d(p))^{2r} \right)^{\frac{1}{2}}
=: \sqrt{S_1 S_2} ,
}
say.
First, we estimate $S_1$. Note that
\[
\rho(n)^v \asymp_v \prod_{\ell|n} \left(1+\frac{v}{\ell}\right)
= \sum_{a|n} \frac{\mu^2(a) \tau_{v}(a) }{a} ,
\]
for any $v\ge0$. Since
\[
\sum_{\substack{a|n \\ a>x}} \frac{\mu^2(a) \tau_v(a)}{a}
\le \frac{1}{x}\sum_{a|n} \mu^2(a)v^{\omega(a)}
= \frac{(v+1)^{\omega(n)}}{x} \ll_{v,\epsilon} \frac{n^\epsilon}{x} ,
\]
we find that
\eq{C-S-e1}{
S_1 &\ll_r \sum_{\substack{ N^-<p <N^+ \\ p=1+jm}}
\left( \sum_{\substack{a|(k,j) \\ a\le k^{1/5} }}\frac{\mu^2(a)\tau_{2|r|}(a)}{a} + O_r(k^{-1/6}) \right)
\left( \sum_{\substack{b|d(p) \\ b\le k^{1/5} }}\frac{\mu^2(b)\tau_{2s}(b)}{b} + O_s(k^{-1/6})\right) \\
&= \sum_{ \substack{ a,b\le k^{1/5} \\ a|k }} \frac{\mu^2(a)\mu^2(b)\tau_{2|r|}(a)\tau_{2s}(b) }{ab}
\sum_{\substack{N^-<p<N^+ \\ p=1+jm \\ a|j,\ b| d(p) }} 1
+O_{r,s}( k^{11/30} ),
}
using the trivial estimate $\#\{N^-<p<N^+:p\equiv1\mod m\}\ll \sqrt{N}/m=\sqrt{k}$.
The innermost sum in the second line of \eqref{C-S-e1} equals
\[
\sum_{\substack{ h \in \mathbb Z/[a,b]\mathbb Z \\ h \equiv 0 \pmod a \\ (h-mk)^2 \equiv 4k \pmod b}}
\sum_{\substack{ N^-<p <N^+ \\ p=1+jm \\ j\equiv h \pmod{[a,b]} }} 1
\ll \frac{\sqrt{N}}{\phi(m[a,b])\log(2k)}
\sum_{\substack{ h \in \mathbb Z/[a,b]\mathbb Z \\ h\equiv 0 \pmod a \\ (h-mk)^2 \equiv 4k \pmod b}} 1
\le \frac{\sqrt{N} \tau(b) }{\phi(m[a,b])\log(2k)} ,
\]
where the first inequality follows from the Brun-Titchmarsch inequality and the second from the fact that $b$ is square-free.
Since $\phi(m[a,b])\ge\phi(m)\phi([a,b])$, relation \eqref{C-S-e1} becomes
\eq{estimation of S_1}{
S_1 &\ll_{r,s} \frac{\sqrt{N}}{\phi(m)\log(2k)} \sum_{ \substack{ a,b\le k^{1/5} \\ a|k }}
\frac{\mu^2(a)\mu^2(b)\tau_{2|r|}(a)\tau_{2s}(b)^2 }{a\cdot b\cdot \phi([a,b])}
+ k^{11/30}
\ll_{r,s} \frac{\sqrt{N}}{\phi(m)\log(2k)} .
}
Next, we turn to the estimation of
\[
S_2 = \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }} \L(k^2d(p))^{2r} .
\]
Our first task is to replace the $L$-values that appear in the above sum with truncated Euler products. We set
\[
S_3 = \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }} \L(k^2d(p);z^{80000})^{2r}
\]
with $z=\log(4k)$ and estimate the error
\[
R:=S_2-S_3
\]
using Lemma~\ref{lemmashortproduct}.
First note that since $d(p)$ is a discriminant and $|d(p)|\le 4k$ for $p\in (N^-,N^+)$,
it follows that
\eq{leg-k^2d(p)}{
\leg{k^2d(p)}{\cdot}
}
is periodic modulo $k|d(p)|\le 4k^2$ and its conductor cannot exceed $|d(p)|\le 4k$.
Thus, we may apply Lemma~\ref{lemmashortproduct} with $\alpha=100$ and $Q=4k$.
Now let $d_1=d_1(p)$ be the discriminant of the quadratic number field $\mathbb Q(\sqrt{d(p)})$, so that the character in \eqref{leg-k^2d(p)} is induced by the primitive character $\leg{d_1}{\cdot}$. If $|d_1| \notin \mathcal{E}_{100}(4k)$, then we can approximate $\L(k^2d(p))^{2r}$ very well by $\L(k^2d(p);z^{80000})^{2r}$.
Otherwise, we write $d(p) = d_1 b^2$ and note that
\[
\L(k^2d(p))^{2r}
\le \rho(kb)^{2|r|} \L(d_1)^{2r}
\ll_r \rho(kb)^{2|r|}\cdot \begin{cases}
(\log|d_1|)^{2r} &\text{if}\ r\ge0,\cr
|d_1|^{1/8} &\text{if}\ r<0,
\end{cases}
\]
the second estimate being a consequence of Siegel's theorem.
In any case, we find that
\[
\L(k^2d(p))^{2r}
\ll_r (\rho(kb))^{2|r|} |d_1|^{1/8} \ll_r (kb |d_1|)^{1/8} \le (k |d(p)|)^{1/8} \le (2k)^{1/4}.
\]
Combining the above, we arrive at the estimate
\als{
R &\ll_{r} \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }} \frac{(\log\log k)^{2|r|} }{\log^{100}(2k)}
+ \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m \\ |d_1| \in\mathcal{E}_{100}(4k)}} k^{1/4}
}
Note that if $p=1+jm$ is such that $|d_1| \in\mathcal{E}_{100}(4k)$, then $d(p)=d_1 b^2$ for some $b\in\mathbb{N}$, or equivalently, $(j-mk)^2-d_1 b^2=4k$.
So for each fixed $d_1$ with $|d_1|\in \mathcal{E}_{100}(4k)$, there are at most $4\tau(4k)\ll k^{1/100}$ admissible values of $j$ (and hence of $p$). Consequently,
\eq{estimation of R}{
R \ll_{r}
\sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }} \frac{(\log\log k)^{2|r|} }{\log^{100}(2k)}
+ k^{1/4} \cdot k^{1/100} \cdot |\mathcal{E}_{100}(4k)|
\ll_{r} \frac{\sqrt{N} }{ \log(2k)\phi(m)},
}
by Lemma\ \ref{lemmashortproduct} and the Brun-Titchmarsch inequality.
Finally, we turn to the estimation of $S_3$. First, note that
\[
\L(k^2d(p);z^{80000})^{2r}
\ll_r \L(k^2d(p);\sqrt{z})^{2r}
\ll_r \prod_{ \substack{ \ell\nmid 2pk \\ 2|r|+1< \ell \le \sqrt{z} }} \left(1+2r\cdot \frac{\leg{d(p)}{\ell} }{\ell}\right) ,
\]
by Mertens' estimate, which immediately implies that
\[
S_3 \ll_r
\sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod{m} }}
\prod_{ \substack{ \ell\nmid 2pk \\ 2|r|+1 < \ell \le \sqrt{z} }}
\left(1+2r\cdot \frac{\leg{d(p)}{\ell} }{\ell}\right) .
\]
We cannot estimate this sum as it is because that would require information about primes in arithmetic progressions that are currently not available.
We refer the reader to~\cite{DS-MEG} for a more detailed discussion about this issue.
Instead, we extend the summation from primes $p$ to integers $n$ with no prime factors $\le k^{1/8}$ and we apply Lemma~\ref{lemma-FI} with $D=k^{1/4}$ and $y=k^{1/8}$. Hence
\eq{S3-S4}{
S_3 \ll_r \sum_{\substack{N^-< n <N^+\\ n \equiv 1\pmod {m} }}(\lambda^+*1)(n)
\prod_{ \substack{2|r|+1 < \ell\le\sqrt{z} \\ \ell\nmid 2nk }} \left(1+2r\cdot \frac{\leg{d(n)}{\ell}}{\ell}\right)=:S_4,
}
by the positivity of the above Euler product. Expanding this product to a sum, opening the convolution $(\lambda^+*1)(n)$, and interchanging the order of summation yields
\als{
S_4 &=
\sum_{\substack{\ell|a\ \Rightarrow\ 2|r|+1<\ell\le \sqrt{z} \\ (a,2k)=1}}\frac{\mu^2(a) \tau_{2r}(a) }{a}
\sum_{\substack{N^-<n<N^+\\ (n,a)=1 \\ n\equiv 1\pmod{m} }}
(\lambda^+*1)(n) \leg{d(n)}{a} \nonumber \\
&= \sum_{\substack{\ell|a\ \Rightarrow\ 2|r|+1<\ell\le \sqrt{z} \\ (a,2k)=1}}\frac{\mu^2(a)\tau_{2r}(a) }{a}
\sum_{\substack{b \leq k^{1/4}\\ (b,am)=1}}\lambda^+(b)
\sum_{\substack{N^-<n<N^+\\ (n,a)=1,\, b|n \\ n\equiv 1\pmod{m} }}\leg{d(n)}{a} .
}
Splitting the integers $n\in(N^-,N^+)$ according to the congruence class of $d(n)\pmod a$, we deduce that
\eq{S_3}{
S_4 = \sum_{\substack{\ell |a\ \Rightarrow\ 2|r|+1<\ell\le \sqrt{z} \\ (a,2k)=1}}
\frac{\mu^2(a)\tau_{2r}(a)}{a}
\sum_{\substack{b \leq k^{1/4}\\ (b,am)=1}}\lambda^+(b)
\sum_{c\in\mathbb Z/a\mathbb Z} \leg{c}{a} S(a,b,c),
}
where
\[
S(a,b,c)
:= \#\left\{ N^-<n<N^+ :
\begin{array}{ll}
n\equiv 1\pmod {m} & (n,a)=1 \\
n\equiv 0\pmod b & d(n)\equiv c\pmod{a}
\end{array}
\right\}.
\]
We fix $a$, $b$ and $c$ as above and calculate $S(a,b,c)$. Set $n=1+j m$, and define $\Delta(j)=(j-mk)^2-4k$, so that $d(n)=\Delta(j)$.
Note that $n$ is counted by $S(a,b,c)$ if and only if $mk - 2\sqrt{k} < j < mk+2\sqrt{k}$, $\Delta(j)\equiv c\pmod a$, $1+jm\equiv0\pmod b$ and $(1 + jm, a) = 1$. Thus we have that
\begin{equation}\label{T}
S(a,b,c)=\left(\frac{4\sqrt{k}}{ab}+O(1)\right)J(a,b,c),
\end{equation}
where
\[
J(a,b,c) :=\#\{j\in \mathbb Z/ab\mathbb Z :
\Delta(j)\equiv c\pmod a,\
1 + jm \equiv 0 \pmod b,\
(1 + jm, a) = 1\}.
\]
By the Chinese remainder theorem, we find that
\[
J(a,b,c)=U(a,c) := \#\{j\in \mathbb Z/a\mathbb Z: \Delta(j)\equiv c\pmod a,\ (1+ jm,a) = 1\},
\]
since $(b,m)=1$, and thus there is exactly one solution modulo $b$ to the equation $1+jm\equiv 0\pmod b$. Note that $U(a,c)\le \tau(a)$ by Lemma \ref{generic quad lemma} and that
\[
\sum_{c\in\mathbb Z/a\mathbb Z}\leg{c}{a} U(a,c) = T(a),
\]
where $T(a)$ is defined by relation \eqref{T(n) def}. Together with relations~\eqref{S_3} and~\eqref{T}, this implies that
\als{
S_4 &= 4\sqrt{k} \sum_{\substack{\ell|a\ \Rightarrow\ 2|r|+1<\ell\le \sqrt{z} \\ (a,2k)=1}}\frac{\mu^2(a)\tau_{2r}(a)T(a)}{a^2}
\sum_{\substack{b \leq k^{1/4}\\ (b,am)=1}}\frac{\lambda^+(b)}b \\
&\quad + O \left( k^{1/4} \sum_{ P^+(a)\le\sqrt{z} } \mu^2(a)\tau_{2|r|}(a)\tau(a) \right) .
}
The error term in the above estimate is
\[
\ll k^{1/4} \sum_{ P^+(a)\le\sqrt{z} } \mu^2(a)\tau_{2|r|}(a)\tau(a)
= k^{1/4} \prod_{\ell\le \sqrt{z}} (1+4|r|) \ll_r k^{1/3} .
\]
Finally, note that $|T(a)|\le \tau(a)$ for square-free values of $a$, by Proposition \ref{odd primes prop}. So applying Lemma~\ref{lemma-FI} we conclude that
\als{
S_4 &\ll_r \sqrt{k} \sum_{\substack{P^+(a)\le\sqrt{z} \\ (a,2k )=1}}\frac{\mu^2(a) \tau(a) \tau_{2|r|}(a)}{a^2}
\prod_{\substack{\ell\le k^{1/8}\\ \ell\nmid am}}\left(1-\frac1{\ell}\right)+k^{1/3} \\
&\ll \sqrt{k}\sum_{\substack{P^+(a)\le\sqrt{z} \\ (a,2k )=1}}\frac{\mu^2(a) \tau(a) \tau_{2|r|}(a)}{a^2}
\frac1{\log(2k)} \frac{m}{\phi(m)} \frac{a}{\phi(a)} +k^{1/3} .
}
Inserting this estimate in \eqref{S3-S4}, we obtain the upper bound
\eq{S_3 e2}{
S_3 \ll_r \frac{\sqrt{k}}{\log(2k)}\frac{m}{\phi(m)}\sum_{(a,2k)=1}\frac{\mu^2(a)\tau(a)\tau_{2|r|}(a)}{a\phi(a)}
\ll_r \frac{\sqrt{k}}{\log(2k)}\frac{m}{\phi(m)}.
}
Combining the above inequality with relations \eqref{C-S}, \eqref{estimation of S_1}, and \eqref{estimation of R} completes the proof of the proposition.
\end{proof}
\section{Approximating $M(G)$}\label{approx}
In this section, we prove Theorem \ref{approximation of M(G)}. We start with a preliminary lemma.
\begin{lma}\label{prime sum} Let $N=m^2k>1$ and $d(p)=d_{m,k}(p)$. If $1\le q\le h \le \sqrt{N}$ and $(a,q)=1$, then
\[
\sum_{\substack{N^-<p\le N^+ \\ p\equiv a\mod q }} \sqrt{|d(p)|}
= \frac{2\pi mk}{\phi(q)\log N}
+ O\left( \frac{h}{\sqrt{N}} \cdot \frac{mk}{q} + \frac{\sqrt{k}}{h\log N} \int_{N^-}^{N^+} E(y,h;q) dy \right) .
\]
\end{lma}
\begin{proof}
We note the trivial bound $\#\{t<p\le t+h:p\equiv a\mod q\}\ll h/q$, which we will use several times throughout the proof. We have that
\eq{prime sum log insertion}{
\sum_{\substack{N^-<p\le N^+ \\ p\equiv a\mod q }} \sqrt{|d(p)|}
= \sum_{\substack{N^-<p\le N^+ \\ p\equiv a\mod q }} \frac{\sqrt{|d(p)|}\log p}{\log N}
+ O\left( \frac{\sqrt{k}}{q} \right) .
}
Note that if $t=N+1+2\sqrt{N}u_0$ and $u_0\in[-1+2\eta,1-\eta]$ with $\eta:=h/\sqrt{4N}$, then
\als{
\sqrt{|d(t)|} = 2\sqrt{k}\cdot \sqrt{1-u_0^2}
&= \frac{2\sqrt{k}}{\eta}\int_{u_0-\eta}^{u_0} \sqrt{1-u^2} \,\mathrm{d} u
+ O\left(\frac{\eta\sqrt{k}}{\sqrt{1-u_0^2}}\right) \\
&=\frac{4mk}{h}\int_{u_0-\eta}^{u_0} \sqrt{1-u^2} \, \mathrm{d} u
+ O\left(\frac{h\sqrt{k}}{\sqrt{4N-(N+1-t)^2}}\right) .
}
Therefore
\als{
\sum_{\substack{N^-<p\le N^+ \\ p\equiv a\mod q }} \frac{\sqrt{|d(p)|}\log p}{\log N}
&= \sum_{\substack{10h+ N^-<p\le -10h+N^+\\ p\equiv a\mod q }} \frac{\sqrt{|d(p)|}\log p}{\log N}
+ O\left( \frac{h^{1/2}N^{1/4}}{m} \cdot \frac{h}{q}\right) \\
&= \frac{4mk}{h\log N}\sum_{\substack{N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }} (\log p)
\int_{\frac{p-N-1-h}{2\sqrt{N}}}^{\frac{p-N-1}{2\sqrt{N}}}
\sqrt{1-u^2} \,\mathrm{d} u \\
&\quad + O\left( \sum_{\substack{N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }}
\frac{h\sqrt{k}}{\sqrt{(N^+-p)(p-N^-)}}
+ \frac{h^{3/2}N^{1/4}}{mq} \right) \\
&= \frac{4mk}{h\log N} \int_{-1+9\eta}^{1-10\eta} \sqrt{1-u^2}
\sum_{\substack{N+1+2u\sqrt{N}<p\le N+1+2u\sqrt{N}+h \\ N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }}
(\log p) \,\mathrm{d} u \\
&\quad + O\left( \sum_{\substack{N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }}
\frac{h\sqrt{k}}{\sqrt{(N^+-p)(p-N^-)}}
+ \frac{h^{3/2}N^{1/4}}{mq} \right) .
}
First, we simplify the main term. If $u\in[-1+10\eta,1-11\eta]$, then the condition that $N^- +10h<p\le N^+ - 10h$ can be discarded. On the other hand, if $u\in[-1,1]\setminus[-1+10\eta,1-11\eta]$, then
\als{
\sqrt{1-u^2}
\sum_{\substack{N+1+2u\sqrt{N}<p\le N+1+2u\sqrt{N}+h \\ N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }}
(\log p)
&\le \sqrt{1-u^2}
\sum_{\substack{N+1+2u\sqrt{N}<p\le N+1+2u\sqrt{N}+h \\ p\equiv a\mod q }}
(\log p) \\
&\ll \sqrt{\eta}\cdot \frac{h\log N}{q} .
}
Therefore
\als{
&\int_{-1+9\eta}^{1-10\eta} \sqrt{1-u^2}
\sum_{\substack{N+1+2u\sqrt{N}<p\le N+1+2u\sqrt{N}+h \\ N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }}
(\log p) \,\mathrm{d} u \\
&\qquad= \int_{-1}^1 \sqrt{1-u^2}
\sum_{\substack{N+1+2u\sqrt{N}<p\le N+1+2u\sqrt{N}+h \\ p\equiv a\mod q }}
(\log p) \, \mathrm{d} u
+ O\left( \frac{\eta^{3/2} h\log N}{q} \right) \\
&\qquad= \int_{-1}^{1} \sqrt{1-u^2} \frac{h}{\phi(q)} \, \mathrm{d} u
+ O\left( \int_{-1}^1 E(N+1+2u\sqrt{N},h;q) \mathrm{d} u + \frac{h^{5/2}\log N}{N^{3/4} q} \right) \\
&\qquad=\frac{\pi}{2}\cdot \frac{h}{\phi(q)} +
O\left( \frac{1}{\sqrt{N}} \int_{N^-}^{N^+} E(y,h;q) \mathrm{d} y + \frac{h^{5/2} \log N}{N^{3/4} q} \right) .
}
Consequently,
\als{
\sum_{\substack{N^-<p\le N^+ \\ p\equiv a\mod q }} \sqrt{|d(p)|}
&= \frac{2\pi mk}{\phi(q)\log N}
+ O\left( \sum_{\substack{N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }} \frac{h\sqrt{k}}{\sqrt{(N^+-p)(p-N^-)}}\right) \\
&+O\left( \frac{\sqrt{k}}{h\log N} \int_{N^-}^{N^+} E(y,h;q) dy + \frac{\sqrt{k}}{q} + \frac{h^{3/2}N^{1/4}}{mq}\right) ,
}
where the term $\sqrt{k}/q$ inside the big-Oh comes from \eqref{prime sum log insertion}.
It remains to bound
\[
\sum_{\substack{N^- +10h<p\le N^+ - 10h \\ p\equiv a\mod q }} \frac{1}{\sqrt{(N^+-p)(p-N^-)}} .
\]
We break this sum into two pieces, according to whether $p\le N+1$ or $p>N+1$. Note that
\als{
\sum_{\substack{N^- +10h<p\le N+1 \\ p\equiv a\mod q }} \frac{1}{\sqrt{(N^+-p)(p-N^-)}}
&\ll N^{-1/4}\sum_{\substack{N^- +10h<n\le N+1 \\ n\equiv a\mod q }} \frac{1}{\sqrt{n-N^-}} .
}
We cover the range of summation by intervals of length $h$ to find that
\als{
\sum_{\substack{N^- +10h<p\le N+1 \\ p\equiv a\mod q }} \frac{1}{\sqrt{(N^+-p)(p-N^-)}}
&\ll N^{-1/4} \sum_{1\le j\le 2\sqrt{N}/h} \frac{1}{\sqrt{jh}} \cdot
\sum_{\substack{N^-+jh<n\le N^-+jh+h \\ n\equiv a\mod q }} 1 \\
&\ll \frac{\sqrt{h}}{N^{1/4}q} \sum_{1\le j\le 2\sqrt{N}/h} \frac{1}{\sqrt{j}}
\ll \frac{1}{q} .
}
Similarly, we find that
\[
\sum_{\substack{N+1<p\le N^+ -10h \\ p\equiv a\mod q }} \frac{1}{\sqrt{(N^+-p)(p-N^-)}} \ll \frac{1}{q}
\]
too, which implies that
\als{
\sum_{\substack{N^-<p\le N^+ \\ p\equiv a\mod q }} \sqrt{|d(p)|}
&= \frac{2\pi mk}{\phi(q)\log N}
+ O\left( \frac{\sqrt{k}}{h\log N} \int_{N^-}^{N^+} E(y,h;q) dy + \frac{h\sqrt{k}}{q} + \frac{h^{3/2}N^{1/4}}{mq}\right) .
}
Since $h^{3/2}=N^{3/4} (h/\sqrt{N})^{3/2}\le N^{3/4} (h/\sqrt{N})$, the lemma follows.
\end{proof}
Using the above result and the results of Section \ref{local computations}, we will prove Theorem \ref{approximation of M(G)}. But first, we need to introduce some additional notation and state another intermediate result. Set
\eq{J_r(v) def}{
J_r(v) = \{1\le j\le 2^{2v+3}:(j-mk)^2\equiv 4k+4^vr\pmod{2^{2v+3}}, \, jm\equiv 0\pmod 2\}
}
and
\eq{J(v)}{
\mathcal J(v) = \frac{1}{2^{v_0-1}} \sum_{r\in\{0,1,4,5\}} \frac{|J_r(v)|}{2-\leg{r}{2}},
\quad\text{where}\quad
v_0 =
\begin{cases}
2 &\text{if}\ 2\nmid m,\cr
3 &\text{if}\ 2|m.
\end{cases}
}
Finally, set
\[
\mathcal J = \sum_{\substack{v\ge0 \\ (2^v,k)=1}} \frac{\mathcal J(v)}{8^v} .
\]
Then we have the following formula.
\begin{lma}\label{formula for J}
\[
\mathcal J =
\begin{cases}
\frac{2}{3} &\mbox{if $2\nmid mk$},\cr
\frac{3}{2} &\mbox{if $2\mid (m,k)$},\cr
1 &\mbox{if $2\mid mk$, $2\nmid (m,k)$}.
\end{cases}
\]
\end{lma}
We postpone the proof of this lemma till the last section.
\begin{proof}[Proof of Theorem \ref{approximation of M(G)}] We will show the theorem with $8\epsilon\in(0,1/3]$ in place of $\epsilon$ and when $k$ is large enough in terms of $\epsilon$, which is clearly sufficient. Our starting point is Lemma \ref{formula for M(G)}, which states that
\[
M(G)
=\sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}
\sum_{\substack{f^2\mid d(p), \, (f,k)=1 \\ d(p)/f^2\equiv 1,0\pmod 4 }}
\frac{\sqrt{|d(p)|} \L(d(p)/f^2) }{2\pi f} ,
\]
where $N=m^2k$ and $d(p)=d_{m,k}(p)=((p-N-1)^2-4N)/m^2$ as usual. If $p=1+jm$, then $d(p)=(j-mk)^2-4k$.
Therefore, if $\ell$ is an odd prime dividing $k$, so that $(\ell,f)=1$ for $f$ as in the above sum, then
\[
\leg{d(p)/f^2}{\ell}
=\leg{d(p)}{\ell} = \leg{j}{\ell}^2 ,
\]
Next, we write $f=2^vg$ with $g$ odd and consider $r\in\{0,1,4,5\}$ such that $d(p)/f^2\equiv r\mod 8$.
Then we have that $\leg{d(p)/f^2}{2}=\leg{r}{2}$. Moreover, since $g^2\equiv1\mod{8}$, we have that
\[
d(p)/f^2
\equiv d(p)/2^{2v}\mod{8},
\]
Therefore, the conditions $f^2|d(p)$ and $d(p)/f^2\equiv r\mod 8$ are equivalent to having $d(p)\equiv 4^v r\mod{2^{2v+3}}$ and $g^2|d(p)$. Setting
\[
\rho(g,d) = \prod_{\ell|g} \left(1-\frac{\leg{d}{\ell}}{\ell} \right)^{-1}
\]
then gives us that
\[
\L(d(p)/f^2)
= \L((2kg)^2d(p)) \frac{\rho(g,d(p)/g^2)}{1-\leg{r}{2}/2} \prod_{\ell|k,\,\ell\nmid 2j}\left(1-\frac{1}{\ell}\right)^{-1}.
\]
Since
\[
\prod_{\ell|k,\,\ell\nmid 2j}\left(1-\frac{1}{\ell}\right)^{-1}
= \sum_{\substack{a|k \\ (a,2j)=1}} \frac{\mu^2(a)}{\phi(a)},
\]
we deduce that
\als{
M(G)
&= \sum_{r\in\{0,1,4,5\}} \frac{1}{2-\leg{r}{2}} \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}
\sum_{\substack{a|k \\ (a,2j)=1}}
\sum_{\substack{v\ge 0,\, (2^v,k)=1 \\ d(p)\equiv 4^v r\mod{2^{2v+3}} }}
\sum_{\substack{g^2\mid d(p)\\ (g,2k)=1 }} \frac{\mu^2(a) \sqrt{|d(p)|}}{\pi 2^v\phi(a)g} \\
&\qquad \times \rho(g,d(p)/g^2) \L((2kg)^2d(p) ) .
}
We now use Lemma~\ref{lemmashortproduct} to replace the $L$-value $\L((2kg)^2d(p))$ by a suitably truncated product.
Arguing as in the proof of relation~\eqref{estimation of R}, we note that $\leg{(2kg)^2d(p)}{\cdot}$ is a character modulo
$2kg|d(p)|\le 16k^{5/2}$ with conductor not exceeding $|d(p)|\le 4k$.
Thus, we may apply Lemma~\ref{lemmashortproduct} with $Q=4k$ and $5\alpha$ in place of $\alpha$ to replace $\L((2kg)^2d(p))$ by $\L((2kg)^2d(p); z)$,
where we take $z=(\log(4k))^{200\alpha^2}$.
The result is that
\als{
M(G)
&= \sum_{r\in\{0,1,4,5\}} \frac{1}{2-\leg{r}{2}} \sum_{\substack{N^-<p<N^+ \\ p=1+jm,\,j\ge1 }}
\sum_{\substack{a|k \\ (a,2j)=1}}
\sum_{\substack{ (2^v,k)=1 \\ d(p)\equiv 4^v r\mod{2^{2v+3}} }}
\sum_{\substack{g^2\mid d(p) \\ (g,2k)=1 }} \frac{\mu^2(a) \sqrt{|d(p)|}}{\pi 2^v\phi(a) g} \\
&\qquad \times \rho(g,d(p)/g^2) \L((2kg)^2d(p) ; z )
+O_{\alpha}\left(\frac{k}{(\log k)^{\alpha}} \right) .
}
Next, we notice that we can truncate the sums over $a,g$ and $v$ at the cost of a small error term.
More precisely, using the crude bound
\[
\rho(g,d(p)/g^2)\L((2kg)^2d(p);z) \ll \frac{g}{\phi(g)} \log(2kg|d(p)|)\ll (\log k)^2 ,
\]
we find that the contribution to $M(G)$ by those summands with $\max\{a,g,2^v\}>k^\epsilon$ is
\eq{divisor bound}{
\ll \frac{\sqrt{k} (\log k)^3}{k^\epsilon} \sum_{\substack{N^-<p<N^+ \\ p \equiv 1 \pmod m }}
\sum_{\substack{a|k \\ (2^vg)^2|d(p)}} 1
\ll_\epsilon k^{(1-\epsilon)/2} \sum_{\substack{ N^-<n<N^+ \\ n\equiv 1\mod m }} 1
\ll k^{1-\epsilon/2}
}
by the bound $\tau(n)\ll_\delta n^\delta$, with $\delta < \epsilon/4$. Moreover,
\[
\L((2kg)^2d(p) ; z)
= \sum_{\substack{P^+(n)\le z \\ (n,2kg)=1}} \frac{\leg{d(p)}{n} }{n}
= \sum_{\substack{P^+(n)\le z,\, n\le k^\epsilon \\ (n,2kg)=1}} \frac{ \leg{d(p)}{n} }{n}
+ O_{\epsilon,\alpha}\left( (\log k)^{-\alpha-10}\right)
\]
by Lemma \ref{smooth}. Therefore,
\als{
M(G) &= \sum_{r\in\{0,1,4,5\}} \frac{1}{2-\leg{r}{2}} \sum_{\substack{a|k,\, a\le k^\epsilon \\ (a,2)=1}}
\sum_{\substack{2^v\le k^\epsilon \\ (2^v,k)=1}}
\sum_{\substack{g\le k^\epsilon \\ (g,2k)=1 }}
\sum_{\substack{P^+(n)\le z,\, n\le k^\epsilon \\ (n,2kg)=1}} \frac{\mu^2(a)}{\pi 2^v\phi(a) g n} \\
&\qquad \times
\sum_{\substack{N^-<p<N^+ \\ p=1+jm,\,j\ge1 \\ (a,j)=1,\, g^2|d(p) \\ d(p)\equiv 4^v r \mod{2^{2v+3}} }}
\rho(g,d(p)/g^2) \leg{d(p)}{n}\sqrt{|d(p)|}
+ O_{\alpha,\epsilon} \left(\frac{k}{(\log k)^{\alpha}}\right) .
}
We note that if $d(p)/g^2\equiv b\mod{g}$, then $\leg{d(p)/g^2}{\ell} = \leg{b}{\ell}$ for all $\ell|g$ and consequently,
$\rho(g,d(p)/g^2)=\rho(g,b)$. So summing over possible choices for $d(p)/g^2\mod g$ and $d(p)\mod n$, we deduce that
\als{
M(G) &= \sum_{r\in\{0,1,4,5\}} \frac{1}{2-\leg{r}{2}} \sum_{\substack{a|k,\, a\le k^\epsilon \\ (a,2)=1}}
\sum_{\substack{2^v\le k^\epsilon \\ (2^v,k)=1}}
\sum_{\substack{g\le k^\epsilon \\ (g,2k)=1 }}
\sum_{\substack{P^+(n)\le z,\, n\le k^\epsilon \\ (n,2kg)=1}} \frac{\mu^2(a)}{\pi 2^v \phi(a) g n } \\
&\qquad \times \sum_{b=1}^g \rho(g,b)
\sum_{c=1}^n\leg{c}{n}
S_r(v,a,g,b,n,c) + O_{\alpha,\epsilon} \left(\frac{k}{(\log k)^{\alpha}}\right) ,
}
where
\[
S_r(v,a,g,b,n,c) : = \sum_{\substack{N^-<p\le N^+ \\ p=1+jm,\,j\ge 1,\, (j,a)=1 \\ d(p)\equiv bg^2\mod{g^3} \\
d(p)\equiv 4^vr \mod{2^{2v+3}},\, d(p)\equiv c\mod{n} }} \sqrt{|d(p)|} .
\]
We write $p=1+jm$ and note that $(1+jm,2agn)=1$ if $k$ is large enough, since $2agn\le 2k^{3\epsilon}\le 2k^{1/8}$ by assumption, and $p>N^-=(m\sqrt{k}-1)^2$. Moreover, with this notation we have that $d(p)=\Delta(j):=(j-mk)^2-4k$. So, if we set
\als{
J_r(v,a,g,b,n,c) =
\left\{ j\mod{2^{2v+3} a g^3n}:
\begin{array}{rl}
\Delta(j) \equiv 4^v r \pmod{2^{2v+3} },& \Delta(j) \equiv b g^2 \pmod{g^3}, \\
\Delta(j)\equiv c\mod{n},& (j,a)=1,\\
(1+jm, agn)=1,& jm\equiv 0\pmod2
\end{array}\right\},
}
then we find that
\[
S_r(v,a,g,b,n,c) = \sum_{j\in J_r(v,a,g,b,n,c)}
\sum_{\substack{N^-<p\le N^+ \\ p\equiv 1+jm\mod{2^{2v+3}ag^3nm}}}
\sqrt{|d(p)|} .
\]
Applying Lemma \ref{prime sum} with $h$ as in the statement of the theorem, we deduce that
\als{
\frac{S_r(v,a,g,b,n,c)}{ |J_r(v,a,g,b,n,c)|}
&= \frac{2\pi mk}{\phi(2^{2v+3}ag^3nm)\log N} \\
&\quad + O\left( \frac{k}{4^vag^3n(\log k)^{\alpha+1}}
+ \frac{\sqrt{k}}{h\log k} \int_{N^-}^{N^+} E(y,h;2^{2v+3}ag^3nm) \mathrm{d} y \right) ,
}
by our assumption that $h\le m\sqrt{k}/(\log k)^{\alpha+1}$ and that $m\le \sqrt{k}$. In order to compute the contribution of the above error term to $M(G)$, we note that
\als{
\sum_{b=1}^g \rho(b,g) \sum_{c=1}^n |J_r(b,v,g,a,n,c)|
&\le \sum_{b=1}^g \sum_{c=1}^n
\sum_{ \substack{ j\mod {2^{2v+3} ag^3n} \\ \Delta(j)\equiv bg^2\pmod{g^3} \\
\Delta(j)\equiv 4^vr \pmod{2^{2v+3}} \\ 2|jm,\ \Delta(j)\equiv c \mod{n} }} \frac{g}{\phi(g)}
= \sum_{ \substack{ j\mod {2^{2v+3} a g^3n} \\ g^2 | \Delta(j),\, 2|jm \\ \Delta(j)\equiv 4^vr\pmod{2^{2v+3}} }} \frac{g}{\phi(g)} \\
&= \frac{g}{\phi(g)} agn \sum_{ \substack{ j\mod {2^{2v+3} g^2} \\ 2|jm,\ g^2 | \Delta(j) \\ \Delta(j)\equiv 4^vr\pmod{2^{2v+3}} }} 1
\ll \frac{ag^2n}{\phi(g)} \cdot \tau(g) \cdot |J_r(v)|
}
by the Chinese remainder theorem and Lemma \ref{generic quad lemma}, where $J_r(v)$ is defined by \eqref{J_r(v) def}. Since we also have that $|J_r(v)|\ll \mathcal J(v) \ll1$ by Lemmas \ref{prime 2 lma1} and \ref{prime 2 lma2} below, we conclude that
\als{
M(G) &= \frac{2mk}{\log N}\sum_{r\in\{0,1,4,5\}} \frac{1}{2-\leg{r}{2}}
\sum_{\substack{a|k,\, a\le k^\epsilon \\ (a,2)=1}}
\sum_{\substack{2^v\le k^\epsilon \\ (2^v,k)=1}}
\sum_{\substack{g\le k^\epsilon \\ (g,2k)=1 }}
\sum_{\substack{P^+(n)\le z,\, n\le k^\epsilon \\ (n,2kg)=1}}
\frac{\mu^2(a)}{2^{3v+v_0}\phi(a)\phi(g^4an^2m)} \\
&\qquad\times \sum_{b=1}^g \rho(g,b)
\sum_{c=1}^n \leg{c}{n}|J_r(b,v,g,a,n,c)|
+ O_{\alpha,\epsilon}\left( \frac{k}{(\log k)^\alpha} + E\right),
}
where $v_0$ is defined by \eqref{J(v)} and
\[
E:= \frac{\sqrt{k}}{h} \sum_{q\le 8k^{7\epsilon}} \tau_3(q) \int_{N^-}^{N^+} E(y,h;mq) \mathrm{d} y ,
\]
since, for any $q\in\mathbb N$, we have that
\[
\sum_{\substack{q=2^{2v+3}ag^3n \\ a|k,\ (a,2)=(gn,2k)=1}} \tau(g) \le \sum_{g|q} \tau(g) = \tau_3(q) .
\]
If we set
\[
I(g,b) = \#\{1\le j\le g^3: \Delta(j)\equiv bg^2\pmod{g^3}, \, (1+jm,g)=1\}
\]
and
\[
F(a) = \#\{ 1\le j\le a: (j,a)=1,\,(1+jm,a)=1\} = \prod_{\ell^w \| a} \ell^{w-1} \left(\ell-1-\leg{m}{\ell}^2\right) ,
\]
then the Chinese remainder theorem implies that
\als{
\sum_{c=1}^n \leg{c}{n} |J_r(v,a,g,b,n,c)|
&= F(a) \cdot |J_r(v)| \cdot I(g,b) \sum_{c=1}^n \leg{c}{n}
\sum_{\substack{ j\mod n \\ \Delta(j)\equiv c\mod{n} \\(1+jm,n)=1 }}1 \\
&= F(a) \cdot |J_r(v)|\cdot I(g,b) \cdot T(n),
}
where $T(n)$ is defined by \eqref{T(n) def}. Therefore,
\[
M(G) = \frac{mk}{\phi(m)\log N} S_1S_2S_3+ O_{\alpha,\epsilon}\left( \frac{k}{(\log k)^\alpha} + E\right),
\]
where
\[
S_1 = \sum_{r\in\{0,1,4,5\}} \frac{2}{2-\leg{r}{2}}
\sum_{\substack{2^v\le k^\epsilon \\ (2^v,k)=1}} \frac{|J_r(v)|}{2^{3v+v_0}}
=\mathcal J + O(k^{-\epsilon}),
\]
by the trivial estimate $|J_r(v)|\ll 4^v$,
\als{
S_2 = \sum_{\substack{a|k,\,a\le k^\epsilon \\(a,2)=1}} \frac{\mu^2(a) F(a)}{\phi(a)a}
\prod_{\ell|a,\,\ell\nmid m} \frac{\ell}{\ell-1}
&= \prod_{\substack{\ell|k\\ \ell\ne 2}} \left( 1+ \frac{\ell-1-\leg{m}{\ell}^2}{(\ell-1)(\ell-\leg{m}{\ell}^2)}\right)
+ O(k^{-\epsilon/2}) \\
&= \prod_{\substack{\ell|k\\ \ell\ne 2}} \frac{\ell^2- \leg{m}{\ell}^2 \ell -1 }{(\ell-1)(\ell-\leg{m}{\ell}^2)} + O(k^{-\epsilon/2})
}
by arguing as in relation \eqref{divisor bound}, and
\[
S_3 = \sum_{\substack{g\le k^\epsilon \\ (g,2k)=1 }} \sum_{b=1}^g \frac{\rho(g,b) I(g,b) S_4(g) }{g^4}
\prod_{\ell|g,\,\ell\nmid m} \frac{\ell}{\ell-1}
\]
with
\[
S_4(g) = \sum_{\substack{P^+(n)\le z,\, n\le k^\epsilon \\ (n,2kg)=1}} \frac{T(n)}{n^2}\prod_{\ell|n,\,\ell\nmid m}\frac{\ell}{\ell-1} .
\]
In the above, to factor $\phi(g^4 a n^2 m)$, we have used the identity
$$
\phi(g^4 a n^2 m) = \phi(m) g^4 a n^2 \prod_{\ell|g,\,\ell\nmid m} \frac{\ell -1}{\ell} \;
\prod_{\ell|a,\,\ell\nmid m} \frac{\ell - 1}{\ell} \;
\prod_{\ell|n,\,\ell\nmid m} \frac{\ell - 1}{\ell}
$$
which holds since $a,n$ and $g$ are pairwise coprime. Note that
\als{
I(g,b) &= \prod_{\ell^w\| g} \#\{j\mod {\ell^{3w}}: (j-mk)^2\equiv 4k+bg^2\mod{\ell^{3w}},\, (1+jm,\ell)=1\} \\
& = \prod_{\ell|g} \left(1+ \leg{(N+1)^2-(4k+bg^2)m^2}{\ell}^2 \leg{4k+bg^2}{\ell}\right) \\
& =\left(1+ \leg{N-1}{\ell}^2\leg{k}{\ell} \right)^{\omega(g)}
}
by Lemma \ref{generic quad lemma}, which is applicable here because $4k+bg^2\equiv 4k\not\equiv 0\mod\ell$ for all primes $\ell|g$. So we see that $I(g,b)$ is independent of $b$, which implies that
\als{
\sum_{b=1}^g \rho(g,b) I(g,b)
&= I(g,0) \prod_{\ell^w\|g} \left( \sum_{b=1}^{\ell^w} \frac{1}{1-\leg{b}{\ell}/\ell} \right) \\
&= I(g,0) \prod_{\ell^w\|g} \left(\ell^{w-1}+ \ell^{w-1}\frac{\ell-1}{2}\frac{1}{1-1/\ell}
+\ell^{w-1}\frac{\ell-1}{2}\frac{1}{1+1/\ell}\right) \\
&= g I(g,0) \prod_{\ell|g} \frac{\ell^2+\ell+1}{\ell(\ell+1)} .
}
Thus we conclude that
\[
S_3 = \sum_{\substack{g\le k^\epsilon \\ (g,2k)=1 }}
\frac{S_4(g)}{g^3} \prod_{\ell|g} \frac{(1+\leg{N-1}{\ell}^2\leg{k}{\ell})(\ell^2+\ell+1)}{(\ell-\leg{m}{\ell}^2)(\ell+1)}.
\]
Moreover, if $P(\ell)$ is as in Corollary \ref{formula for P(ell)}, then we have that
\[
S_4(g) = \frac{P}{\prod_{\ell|g} P(\ell)} \left( 1
+ O\left(\frac{1}{(\log k)^{\alpha+1}}\right) \right) ,
\quad\text{where}\quad
P : = \prod_{\ell\nmid 2k} P(\ell) .
\]
Therefore
\als{
S_3\left( 1+ O\left(\frac{1}{(\log k)^{\alpha+1}}\right) \right)
&= P \cdot \prod_{\ell\nmid2k} \left( 1 + \sum_{w\ge1} \frac{(1+\leg{N-1}{\ell}^2\leg{k}{\ell})(\ell^2+\ell+1)}{\ell^{3w}(\ell-\leg{m}{\ell}^2)(\ell+1)P(\ell)}\right) \\
&= \prod_{\ell\nmid2k} \left( P(\ell) + \frac{1+\leg{N-1}{\ell}^2\leg{k}{\ell}}{(\ell^{2}-1)(\ell-\leg{m}{\ell}^2)} \right) \\
&= \prod_{\ell\nmid2k} \frac{\ell^3 - \leg{m}{\ell}^2 \ell^2 - (1+\leg{m(N-1)}{\ell}^2)\ell }{(\ell^{2}-1)(\ell-\leg{m}{\ell}^2)}\\
&= \prod_{\ell\nmid2N} \left( 1- \frac{\ell \leg{N-1}{\ell}^2+1}{(\ell^{2}-1)(\ell-1)} \right) .
}
Consequently,
\als{
M(G) &= \frac{\mathcal J mk}{\phi(m)\log N} \prod_{\ell\nmid2N} \left( 1- \frac{\ell \leg{N-1}{\ell}^2+1}{(\ell^{2}-1)(\ell-1)} \right)
\prod_{\substack{\ell|k \\ \ell>2}} \left( 1+ \frac{\ell-1-\leg{m}{\ell}^2}{(\ell-1)(\ell-\leg{m}{\ell}^2)}\right) \\
&\quad + O_{\alpha,\epsilon}\left( \frac{k}{(\log k)^\alpha} + E\right) .
}
So the theorem follows by the above estimates together with Lemmas \ref{Aut(G)} and \ref{formula for J}.
\end{proof}
\section{Powers of 2}\label{2}
The goal of this section is to show Lemma \ref{formula for J}, which gives the value of
\begin{equation*}
\mathcal J = \sum_{\substack{v\ge0 \\ (2^v,k)=1}} \frac{\mathcal J(v)}{8^v},
\end{equation*}
where
\begin{equation*}
\mathcal J(v) = \frac{1}{2^{v_0-1}} \sum_{r\in\{0,1,4,5\}} \frac{|J_r(v)|}{2-\leg{r}{2}},
\quad\quad
v_0 =
\begin{cases}
2 &\text{if}\ 2\nmid m,\cr
3 &\text{if}\ 2|m,
\end{cases}
\end{equation*}
and
\begin{equation*}
J_r(v) = \{1\le j\le 2^{2v+3}:(j-mk)^2\equiv 4k+4^vr\pmod{2^{2v+3}}, \, jm\equiv 0\pmod 2\}.
\end{equation*}
We start with the following standard lemma.
\begin{lma}\label{generic quad lemma - prime 2}
We have that
\[
\#\{j\in\mathbb Z/8\mathbb Z : j^2\equiv d\pmod{8}\} =
\begin{cases}
2 &\text{if}\ d\equiv 0,4\mod 8,\\
4 &\text{if}\ d\equiv 1\mod 8,\\
0 &\text{otherwise}.
\end{cases}
\]
Moreover, if $d$ is odd and $e\ge3$, then
\[
\#\{j\in\mathbb Z/2^e\mathbb Z : j^2\equiv d\pmod{2^e}\} =
\begin{cases}
4 &\text{if}\ d\equiv 1\mod 8, \\
0 &\text{otherwise}.
\end{cases}
\]
\end{lma}
We shall use the above lemma to calculate $|J_r(v)|$ and $\mathcal J(v)$ when $(2^v,k)=1$.
First, we note that if $v\ge1$, then $k$ must be odd and
\eq{J_r(v) alt}{
|J_r(v)| = \begin{cases}
2\cdot \#\{ j\mod{ 2^{2v+1} } : j^2\equiv k+4^{v-1} r\mod{2^{2v+1}} \} &\text{if}\ 2|m,\\
0 &\text{if}\ 2\nmid m.
\end{cases}
}
Indeed, when $v\ge1$, the relation $(j-mk)^2\equiv 4k+4^vr\mod{2^{2v+3}}$ implies that $2|(j-mk)$. Since $k$ is odd and we also have that $jm\equiv 0\mod{2}$, we deduce that $2\mid (m,j)$.
Hence, $|J_r(v)|=0$ when $2\nmid m$.
Assuming that $2\mid m$, we write $j=mk+2j'$ and find that
\als{
|J_r(v)| &= \#\{j'\mod{2^{2v+2}} : j'^2\equiv k+4^{v-1} r\mod{2^{2v+1}} \} \\
&= 2\cdot \#\{j\mod{2^{2v+1} } : j^2\equiv k+4^{v-1} r\mod{2^{2v+1}} \} ,
}
as claimed.
\begin{lma}\label{prime 2 lma1} Let $v\ge0$ with $(2^v,k)=1$. If $m$ is odd, then
\[
\mathcal J(v) =
\begin{cases}
1 &\mbox{if $v=0$ and $2|k$},\\
\frac{2}{3} &\mbox{if $v=0$ and $2\nmid k$},\\
0 &\text{if $v\ge1$ and $2\nmid k$}.
\end{cases}
\]
\end{lma}
\begin{proof} The case $v\ge1$ follows by \eqref{J_r(v) alt}. Assume now that $v=0$. Since $m$ is odd, the condition $jm\equiv 0\pmod 2$ implies that every $j\in J_r(v)$ is even. Writing $j=2j'$, we deduce that
\[
|J_r(0)| = \#\{j'\mod 4: (2j'-mk)^2\equiv 4k+r\mod{8}\}
\]
If $k$ is odd, then we must have that $(2j'-mk)^2-4k\equiv -3\mod8$ and thus $r=5$, in which case $|J_r(0)|=4$; otherwise $|J_r(v)|=0$. So
\[
\mathcal J(0) = \frac{1}{2} \cdot \frac{4}{2-(-1)} = \frac{2}{3} .
\]
Finally, assume that $k$ is even. Writing $z=j'-mk/2$, our task reduces to counting solutions to $4z^2\equiv r\mod 8$ with $1\le z\le 4$. If $r\in\{1,5\}$, then there are no such solutions, whereas if $r\in\{0,4\}$, then there are precisely two such solutions. Consequently, when $m$ is odd and $k$ is even,
\[
\mathcal J(0)=\frac{1}{2} \left( \frac{2}{2-0} + \frac{2}{2-0} \right) = 1 ,
\]
and the lemma follows in this case too.
\end{proof}
\begin{lma}\label{prime 2 lma2} Let $v\ge0$ with $(2^v,k)=1$, and suppose that $2|m$.
If $2|k$, then
\[
\mathcal J(0) = \frac{3}{2} .
\]
If $k\equiv 1\pmod 8$, then
\[
\mathcal J(v) = \begin{cases}
\frac{5}{6} &\mbox{if $v=0$}, \\
1 &\mbox{if $v=1$}, \\
2 &\mbox{if $v=2$}, \\
\frac{14}{3} &\mbox{if $v\ge3$}.
\end{cases}
\]
If $k\equiv 3,7\pmod 8$, then
\[
\mathcal J(v) = \begin{cases}
\frac{5}{6} &\mbox{if $v=0$}, \\
\frac{4}{3} &\mbox{if $v=1$}, \\
0 &\mbox{if $v\ge2$}.
\end{cases}
\]
If $k\equiv 5\pmod 8$, then
\[
\mathcal J(v) = \begin{cases}
\frac{5}{6} &\mbox{if $v=0$}, \\
1 &\mbox{if $v=1$}, \\
\frac{8}{3} &\mbox{if $v=2$}, \\
0 &\mbox{if $v\ge3$}.
\end{cases}
\]
\end{lma}
\begin{proof}
First, we calculate $|J_r(0)|$. Note that the condition $jm\equiv 0\mod2$ is trivially satisfied now since $2|m$.
Therefore, a change of variable and Lemma \ref{generic quad lemma - prime 2} imply that
\eq{J_r(0)}{
|J_r(0)| = \#\{j\mod 8: j^2\equiv 4k+r\mod{8}\} =
\begin{cases}
2 &\text{if}\ 4k+r\equiv 0,4\mod 8,\\
4 &\text{if}\ 4k+r\equiv 1\mod 8,\\
0 &\text{if}\ 4k+r\equiv 5\mod 8 .
\end{cases}
}
Thus,
\[
\mathcal J(0) = \begin{cases}
\frac{1}{4}\left( \frac{2}{2-0}+\frac{4}{2-1} + \frac{2}{2-0} + \frac{0}{2-(-1)} \right) = \frac{3}{2}
&\text{if}\ 2|k,\\
\frac{1}{4} \left(\frac{2}{2-0}+\frac{0}{2-1} + \frac{2}{2-0}+ \frac{4}{2-(-1)}\right) = \frac{5}{6}
&\text{if}\ 2\nmid k .
\end{cases}
\]
Next assume that $v\ge1$, and note that the condition $(2^v,k)=1$ means that we only need consider this case when $k$ is odd.
By relation \eqref{J_r(v) alt}, we have that
\[
|J_r(v)| = 2\cdot \#\{ j\mod{2^{2v+1}}: j^2\equiv k+4^{v-1} r\mod{2^{2v+1}} \}.
\]
Now if $v\ge2$, then Lemma~\ref{generic quad lemma - prime 2} implies that $|J_r(v)|=2\cdot 4=8$ or $|J_r(v)|=0$
according to whether $k+4^{v-1}r\equiv 1\mod{8}$ or not.
Therefore, when $v\ge2$,
\[
\mathcal J(v) = \begin{cases}
\frac{1}{4} \left(\frac{8}{2-0}+\frac{8}{2-0}\right) = 2
&\text{if}\ v=2\ \text{and}\ k\equiv 1\mod 8,\\
\frac{1}{4} \left( \frac{8}{2-1} + \frac{8}{2-(-1)}\right) = \frac{8}{3}
&\text{if}\ v=2\ \text{and}\ k\equiv 5\mod 8,\\
\frac{1}{4} \left(\frac{8}{2-0}+\frac{8}{2-1} + \frac{8}{2-0}+ \frac{8}{2-(-1)}\right) = \frac{14}{3}
&\text{if}\ v\ge3\ \text{and}\ k\equiv 1\mod 8,\\
0 &\text{otherwise} .
\end{cases}
\]
Finally, we consider the case $v=1$.
Using Lemma~\ref{generic quad lemma - prime 2} again, we have
\[
|J_r(1)| = 2\cdot \#\{ j\mod{8}: j^2\equiv k+r\mod 8 \}
= \begin{cases}
4 &\text{if}\ k+r\equiv 0,4\mod 8,\\
8 &\text{if}\ k+r\equiv 1\mod 8,\\
0 &\text{otherwise}.
\end{cases}
\]
Therefore,
\[
\mathcal J(1) = \begin{cases}
\frac{1}{4} \cdot \frac{8}{2-0} = 1
&\text{if}\ k\equiv 1,5\mod 8,\\
\frac{1}{4} \left( \frac{4}{2-1} + \frac{4}{2-(-1)} \right) = \frac{4}{3}
&\text{if}\ k\equiv 3,7 \mod 8 ,
\end{cases}
\]
which completes the proof of the lemma.
\end{proof}
Lemma \ref{formula for J} now follows as a direct consequence of Lemmas \ref{prime 2 lma1} and \ref{prime 2 lma2}.
|
\section{introduction}\label{intro}
Scale invariance -- also Weyl gauge invariance -- is being, once again, the subject of intensive debate \cite{bars, bars', bars'', kallosh-linde, carrasco-kallosh, bars-ult, padilla, percacci, prester, shaposhnikov, scholz, indios, quiros-2000, quiros-2013, quiros-2014}. It was Weyl \cite{weyl} who made the first serious attempt to create a scale-invariant theory of gravity (and of electromagnetism). Due to an unobserved broadening of the atomic spectral lines this attempt had a very short history \cite{perlick, novello, scholz-h}. Resurrection of scale invariance is associated with the following prototype action \cite{deser} (see also the latter \cite{smolin}):
\bea S=\int d^4x\sqrt{|g|}\left[\frac{\phi^2}{12}\,R+\frac{1}{2}(\der\phi)^2+\frac{\lambda}{12}\,\phi^4\right],\label{deser-action}\eea where $\phi$ is the dilaton\footnote{The role of the dilaton $\phi$ is to modify the local strength of gravity through the effective gravitational coupling $G(\phi)=(3/4\pi)\phi^{-2}$.} and $(\der\phi)^2\equiv g^{\mu\nu}\der_\mu\phi\der_\nu\phi$. Since, under the Weyl gauge transformations\footnote{In this paper we shall use interchangeably the terms Weyl gauge transformations and (local) scale transformations.}
\bea g_{\mu\nu}\rightarrow\Omega^{-2}g_{\mu\nu},\;\phi\rightarrow\Omega\,\phi,\label{scale-t}\eea where the non-vanishing smooth function $\Omega^2=\Omega^2(x)$ is the conformal factor, the combination $\sqrt{|g|}[\phi^2R+6(\der\phi)^2]$ is kept unchanged -- as well as the scalar density $\sqrt{|g|}\phi^4$ -- then the action (\ref{deser-action}) is invariant under (\ref{scale-t}). Any scalar field which appears in the gravitational action the way $\phi$ does, is said to be conformally coupled to gravity. Hence, for instance, the following action \cite{bars, bars', bars'', bars-ult, prester}:
\bea S=\int d^4x\sqrt{|g|}\left[\frac{\left(\phi^2-\sigma^2\right)}{12}\,R+\frac{1}{2}(\der\phi)^2-\frac{1}{2}(\der\sigma)^2\right],\label{bars-action}\eea is also invariant under (\ref{scale-t}) since both $\phi$ and $\sigma$ are conformally coupled to gravity, provided that the additional scalar field $\sigma$ transforms in the same way as $\phi$: $\sigma\rightarrow\Omega\,\sigma$. For the coupling $\propto (\phi^2-\sigma^2)^{-1}$ to be positive and the theory Weyl-invariant, the scalar $\vphi$ must have a wrong sign kinetic energy -- just like in (\ref{deser-action}) -- potentially making it a ghost. However, the local Weyl gauge symmetry compensates, thus ensuring the theory is unitary \cite{bars, bars', bars''}.
In general any theory of gravity can be made Weyl-invariant by introducing a dilaton. In \cite{percacci} it is shown how to construct renormalization group equations for such kind of theories, while in \cite{odintsov} it has been shown that scale invariance is very much related with the effect of asymptotic conformal invariance, where quantum field theory predicts that theory becomes effectively conformal invariant. In Ref. \cite{padilla} the authors present the most general actions of a single scalar field and two scalar fields coupled to gravity, consistent with second order field equations in four dimensions (4D), possessing local scale invariance.\footnote{I want to underline that, whenever I cite \cite{padilla} in the present work as an example of papers where the geometrical aspect of the gravitational theories is not explored in due details, it should be recognized that the aim of the authors of that paper was to construct scale invariant actions without making emphasis in the geometrical aspect of the corresponding theories.} It has been shown that Weyl-invariant dilaton gravity provides a description of black holes without classical spacetime singularities \cite{prester}. Singularities appear due to ill-behavior of gauge fixing conditions, one example being the gauge in which theory is classically equivalent to standard General Relativity (GR). In \cite{bars} (see also \cite{bars', bars''}) the authors show how to lift a generic non-scale invariant action in Einstein frame into a Weyl-invariant theory and present a new general form for Lagrangians consistent with Weyl symmetry. Advantages of such a conformally invariant formulation of particle physics and gravity -- claim the authors -- include the possibility of constructing geodesically complete cosmologies. In this regard see critical comments in \cite{carrasco-kallosh} and the reply \cite{bars-ult}.
A noticeable deficiency of several works on scale invariance -- among them those of references \cite{bars, bars', bars'', kallosh-linde, carrasco-kallosh, bars-ult, padilla, percacci, prester, shaposhnikov, deser} -- is the lack of any kind of discussion about whether the Weyl symmetry of the action is shared by the geometrical laws which govern the affine structure of the underlying manifold. That this poses an actual problem for theories claimed to be Weyl-invariant will be evident from the discussion in this paper. Recall that any scalar-tensor theory -- this includes the effective string theory \cite{wands} -- being a metric theory of gravity, is deeply connected with the geometric structure of the spacetime manifold. Besides the lack of discussion on the geometrical structure of given theories, in these works it is not discussed how the matter degrees of freedom, other than the conformally coupled scalars and radiation, impact on the scale invariance of a given theory. As we shall show this is a critical issue when discussing on Weyl invariance.
Aim of the present work is to fill the existing gap in the understanding of scale invariance by focusing in geometrical as well as in physical aspects of the issue. It will be shown, in particular, that several theories claimed to be Weyl-invariant do not admit matter fields other than radiation to couple minimally to the dilaton. Only a convenient non-minimal coupling of matter is allowed, which means that these matter degrees of freedom necessarily suffer the effects of an additional five-force (5-force) of non-gravitational origin thus destroying any existing Weyl invariance. As a matter of fact the theories whose action is Weyl gauge invariant but which are associated with Riemannian spacetime manifolds can not be actually scale-invariant since the affine properties of the Riemann space are modified by the scale transformations (\ref{scale-t}). This is reflected in that the equations of motion of time-like test particles, as well as the equations of motion of matter fluxes (continuity equations), are not scale-invariant. In this connection we shall show that, as anticipated by Dicke \cite{dicke}, generalizations of Riemann geometry -- notably Weyl(integrable) geometry \cite{quiros-2000, quiros-2013, quiros-2014, cheng, bib-weyl, books} -- seem to be a better suited arena where to play the scale-invariant laws of gravity.
To start with in the next section we shall explicitly show the very well known -- yet usually forgotten -- fact that the gravitational action (\ref{deser-action}) (the same for (\ref{bars-action})) is nothing but plain Brans-Dicke (BD) theory \cite{bd} with the anomalous value $\omega=-3/2$ of the BD coupling. As it is also well-known, this means that only traceless matter (radiation) can be consistently treated in this theory. Other matter degrees of freedom can not be included in the action in a consistent way unless an appropriate non-minimal coupling with the dilaton is allowed. In sections \ref{riem} and \ref{wing} we explore the geometrical aspect of the Weyl gauge symmetry, an issue which is usually avoided when scale invariance is investigated. In section \ref{riem} we shall show that even the vacuum theory (\ref{deser-action}) (also (\ref{bars-action})) is not actually a scale-invariant setup. This is due to the fact that the equation of motion of time-like point-particles (properly the geodesic curves of the Riemann geometry) are not Weyl-invariant. Other aspects of scale-invariance such as the arising gauge freedom and the possibility of avoidance of several spacetime singularities are also discussed in this section. In section \ref{wing} we investigate the impact of a slight modification of Riemann geometry known as Weyl-integrable geometry (WING) in the study of scale invariance. Given that the laws of Weyl geometry, including the affine properties of space, are invariant under the scale transformations (\ref{scale-t}), it is not surprising that WING can be a natural arena where to investigate the Weyl invariance of the gravitational laws. It will be shown that, if associate a given scale-invariant gravitational action with WING spacetime manifolds, the resulting theory of gravity is actually Weyl-invariant. This result is independent of the matter content of the spacetime. The very debated issue of the possibility to avoid several spacetime singularities just by going into a different gauge will be discussed in section \ref{sing} by means of the study of several gauge-free curvature invariants. Physical discussion of the results and brief conclusions will be given in section \ref{conclu}.
In this paper, for sake of simplicity, we shall focus in theories whose gravitational sector is depicted by the prototype action (\ref{deser-action}). It is evident that the obtained results are safely applicable to any theory where the dilaton and perhaps other scalar fields are conformally coupled to the curvature as, for instance, in (\ref{bars-action}) (see the references \cite{bars, bars', bars'', bars-ult, padilla, prester}). Our analysis is fully classical so that any possible influence of quantum aspects on scale invariance is not considered here.
\section{Anomalous Brans-Dicke coupling}\label{anomal}
A first step towards a precise understanding of the meaning of scale invariance in theories where the gravitational sector is described by the action (\ref{deser-action}) is to realize that the above theory is nothing but plain BD theory \cite{bd} with the anomalous value of the BD coupling parameter $\omega=-3/2$. Actually, under the replacement $$\vphi\rightarrow\frac{\phi^2}{12}\;\Rightarrow\;\sqrt\frac{3}{\vphi}\,\der_\mu\vphi\rightarrow\der_\mu\phi,$$ the action (\ref{deser-action}) is written as (here we omit the quartic potential term) $$S\rightarrow S_\text{BD}=\int d^4x\sqrt{|g|}\left[\vphi\,R+\frac{3}{2}\frac{(\der\vphi)^2}{\vphi}\right].$$
To understand what this entails, let us consider the action (\ref{deser-action}) with the addition of the following matter piece:
\bea S_\text{mat}=\int d^4x\sqrt{|g|}\,{\cal L}_\text{mat}[\chi,\nabla\chi,g],\label{matter-action}\eea where ${\cal L}_\text{mat}$ is the matter Lagrangian and $\chi$ collectively stands for the matter degrees of freedom other than the conformally coupled scalars.\footnote{It is assumed that under (\ref{scale-t}) (see, for instance, the appendix A of Ref. \cite{wands}) $${\cal L}_\text{mat}[\chi,\nabla\chi,g]\rightarrow\Omega^4{\cal L}_\text{mat}[\chi,\nabla\chi,\Omega^{-2}g],$$ where it is explicit that $\sqrt{|g|}\,{\cal L}_\text{mat}$ is unchanged by the Weyl gauge transformations (\ref{scale-t}). Yet, implicitly it is seen that the particles couple to the conformal metric, which results in that these do not follow geodesics (see the discussion below).} The field equations which can be derived from the resulting total action
\bea S_\text{tot}=\int d^4x\sqrt{|g|}\left[\frac{\phi^2}{12}\,R+\frac{1}{2}(\der\phi)^2+\frac{\lambda}{12}\,\phi^4+{\cal L}_\text{mat}\right],\label{tot-action}\eea by taking variations with respect to $g_{\mu\nu}$ and $\phi$ which vanish on the boundary, are the following Einstein equations plus the Klein-Gordon (KG) equation for the dilaton $\phi$ respectively [$G_{\mu\nu}\equiv R_{\mu\nu}-g_{\mu\nu}R/2$]:
\bea &&G_{\mu\nu}=\frac{6}{\phi^2}\,T^\text{mat}_{\mu\nu}-\frac{4}{\phi^2}\left[\der_\mu\phi\der_\nu\phi-\frac{1}{4}g_{\mu\nu}(\der\phi)^2\right]\nonumber\\
&&\;\;\;\;\;\;\;\;\;\;\;\;\;\;+\frac{2}{\phi}\left[\nabla_\mu\nabla_\nu\phi-g_{\mu\nu}\Box\phi\right]+\frac{\lambda}{2}g_{\mu\nu}\phi^2,\label{feq}\\
&&\Box\phi-\frac{1}{6}\,R\phi=\frac{\lambda}{3}\,\phi^3,\label{kg-eq}\eea where $\Box\phi\equiv g^{\mu\nu}\nabla_\mu\nabla_\nu\phi$, and $$T^\text{mat}_{\mu\nu}=-\frac{2}{\sqrt{|g|}}\frac{\der\left(\sqrt{|g|}{\cal L}_\text{mat}\right)}{\der g^{\mu\nu}},$$ is the stress-energy tensor of matter. The trace of (\ref{feq}) $$\Box\phi-\frac{1}{6}\,R\phi=\frac{1}{\phi}\,T^\text{mat}+\frac{\lambda}{3}\,\phi^3,$$ when compared with the KG equation (\ref{kg-eq}) yields to the constraint $T^\text{mat}\equiv g^{\mu\nu}T^\text{mat}_{\mu\nu}=0$. Hence, only traceless matter (radiation) is supported by the field equations (\ref{feq}), (\ref{kg-eq}), which are derived from the action (\ref{tot-action}). This is due to the fact that the resulting theory is just Brans-Dicke theory with the anomalous value $\omega=-3/2$ of the BD coupling parameter.
Let us to forget for a while about the above crucial deficiency of the theory (\ref{tot-action}), and assume anyway that matter degrees of freedom other than radiation and the conformally coupled scalars can be consistently included\footnote{The only possibility is to look for a convenient coupling of the matter Lagrangian to the dilaton.} in (\ref{feq}), (\ref{kg-eq}). The equation of motion of the matter degrees of freedom, properly the conservation of energy and stresses:
\bea \nabla^\kappa T^\text{mat}_{\kappa\mu}=0,\label{matt-cons}\eea is indeed transformed by (\ref{scale-t}). In other words, the equations of motion of matter (with non-vanishing trace) are not Weyl-invariant. Actually, under the conformal transformation of the metric \cite{faraoni-rev}: $g_{\mu\nu}\rightarrow\Omega^{-2}g_{\mu\nu}\;\Rightarrow$
\bea \nabla^\kappa T^\text{mat}_{\kappa\mu}=0\;\rightarrow\;\nabla^\kappa T^\text{mat}_{\kappa\mu}=-\frac{\der_\mu\Omega}{\Omega}\,T^\text{mat}.\label{cons-scale-t}\eea
It is seen that the presence of matter with non-vanishing trace spoils any initially existing scale invariance. In this regard, the claim in \cite{bars''} -- to cite an example, but see also \cite{bars, bars', bars-ult} -- that the theory given by the following Lagrangian (Eq. (1) of Ref. \cite{bars''}):
\bea &&{\cal L}_\text{tot}=\frac{\left(\phi^2-\sigma^2\right)}{12}\,R+\frac{1}{2}(\der\phi)^2-\frac{1}{2}(\der\sigma)^2\nonumber\\
&&\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;-\phi^4f(\sigma/\phi)+{\cal L}_\text{rad}+{\cal L}_\text{mat},\nonumber\eea is Weyl-invariant, is actually misleading. While the above claim were justified if remove the matter term ${\cal L}_\text{mat}$, in general it is incorrect. In \cite{bars} the argument used was that all particles masses $m$ are generated by the Higgs field -- in our notation it is the scalar field $\sigma$ corresponding to the non-vanishing component of the Higgs multiplet in the unitary gauge -- so that $m\propto\sigma(x)$. The authors conclude that due to this fact, since under (\ref{scale-t}) and $\sigma\rightarrow\Omega\,\sigma$; $m\rightarrow\Omega\,m$, $ds\rightarrow\Omega^{-1}ds$, the particle action $S=\int m\,ds$ is Weyl invariant. However, as we shall show in the next section, it is not enough that the action be explicitly scale invariant. Notice, additionally, that if remove the non-minimal coupling of the Higgs field to the curvature [$\sigma^2R\rightarrow 0$], the term $\propto-\phi^4f(\sigma/\phi)$, and perhaps other potential terms like the one $\propto(|\sigma|^2-v^2_0\phi^2)^2$, are invariant under (\ref{scale-t}) even if the resulting total action itself is not scale-invariant. In other words, the conformal coupling of $\sigma$ to the curvature $\sigma^2R+6(\der\sigma)^2$, which includes the non-minimal coupling $\propto \sigma^2R$, is cornerstone to allow for Weyl invariance. This means that even if the mass of particles in ${\cal L}_\text{mat}$ is proportional to the Higgs scalar $\sigma$ and the corresponding classical mechanics Lagrangian is Weyl invariant, a minimal coupling of matter to gravitation breaks the scale invariance of the theory.
In the next section, by means of geometric arguments, we shall show that even in the absence of matter the theory (\ref{deser-action}) -- the same for (\ref{bars-action}) -- is not actually a Weyl invariant setup.
\section{(pseudo)Riemann spacetimes}\label{riem}
Up to this point we have forgotten about the unavoidable discussion on the geometrical aspects of given gravitational theories being metric theories of gravity. In the present and subsequent sections we shall pay due attention to the geometrical aspects of the scale-invariant theories explored here.
Although in the works of \cite{bars, bars', bars'', kallosh-linde, carrasco-kallosh, bars-ult, padilla, prester} no single word is said about the affine geometric structure of the spacetime manifold when discussing on conformal invariance, it is implicitly assumed that it is Riemann geometric theory which dictates the properties of the spacetime just like in general relativity. Likewise, in the present section we shall assume Riemannian spacetimes depicted by the pair $({\cal M},g_{\mu\nu})$, where ${\cal M}$ is the 4D manifold, and $g_{\mu\nu}$ the spacetime metric which obeys the Riemannian metricity condition
\bea \nabla_\mu g_{\nu\lambda}=0,\label{riem-metricity-c}\eea with the covariant derivative operator $\nabla_\mu$ defined in terms of $\{^\mu_{\nu\lambda}\}$ - the Christoffel symbols of the metric.
Here we shall assume that the laws of gravity are governed by the vacuum action (\ref{deser-action}), i. e., no matter degrees of freedom other than the dilaton -- and perhaps other conformally coupled scalar fields like in (\ref{bars-action}) -- are considered. Under the conformal transformation of the metric in (\ref{scale-t}): $g_{\mu\nu}\rightarrow\Omega^{-2}g_{\mu\nu}$, the Christoffel symbols transform like \cite{faraoni-rev, faraoni}
\bea \{^\alpha_{\beta\sigma}\}\rightarrow\{^\alpha_{\beta\sigma}\}-\frac{1}{\Omega}\left(\delta^\alpha_\beta\der_\sigma\Omega+\delta^\alpha_\sigma\der_\beta\Omega-g_{\beta\sigma}\der^\alpha\Omega\right).\label{1-vp}\eea In this case it is understood that, given that Riemann geometry governs the affine properties of the original spacetime, the affine structure of the conformal space is also Riemannian, i. e., the Riemannian metricity condition $$\nabla_\mu g_{\nu\lambda}=0\;\rightarrow\;\nabla_\mu g_{\nu\lambda}=0,$$ is preserved by the conformal transformation of the metric, so that $({\cal M},g_{\mu\nu})\rightarrow({\cal M},g_{\mu\nu}).$
The point to notice is that, while the laws of gravity represented by (\ref{deser-action}) are unchanged by (\ref{scale-t}), the time-like Riemannian geodesics
\bea \frac{d^2x^\mu}{ds^2}+\{^\mu_{\kappa\nu}\}\frac{dx^\kappa}{ds}\frac{dx^\nu}{ds}=0,\label{riem-geod}\eea i. e., the motion equations for time-like point-particles in a curved spacetime $({\cal M},g_{\mu\nu})$ which solves (\ref{deser-action}), are mapped into non-geodesics of the conformal (also Riemannian) space $({\cal M},g_{\mu\nu})$:
\bea \frac{d^2x^\mu}{ds^2}+\{^\mu_{\kappa\nu}\}\frac{dx^\kappa}{ds}\frac{dx^\nu}{ds}=\frac{\der_\kappa\Omega}{\Omega}\frac{dx^\kappa}{ds}\frac{dx^\mu}{ds}-\frac{\der^\mu\Omega}{\Omega},\label{non-geod}\eea where under a convenient re-parametrization, $ds\rightarrow\Omega^{-1}d\tau$, the first term in the right-hand-side (RHS) above can be removed, but the second term $\propto-\Omega^{-1}\der^\mu\Omega$ can not be eliminated at all. Hence, given that (\ref{non-geod}) does not admit an affine parametrization whatsoever, this is not a Riemannian geodesic equation \cite{quiros-2013, wald}.
We are faced with the following situation: while in the original Riemannian spacetime $({\cal M},g_{\mu\nu})$ the time-like point-particles follow Riemannian geodesics (\ref{riem-geod}), in the conformal -- also Riemannian -- spacetime $({\cal M},g_{\mu\nu})$, the free-falling point-particles follow time-like curves which solve (\ref{non-geod}) which are not Riemannian geodesics. In both cases the laws of vacuum gravity are the same:
\bea &&G_{\mu\nu}=\frac{\lambda}{2}g_{\mu\nu}\phi^2-\frac{4}{\phi^2}\left[\der_\mu\phi\der_\nu\phi-\frac{1}{4}g_{\mu\nu}(\der\phi)^2\right]\nonumber\\
&&\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+\frac{2}{\phi}\left[\nabla_\mu\nabla_\nu\phi-g_{\mu\nu}\Box\phi\right],\label{vac-feq}\\
&&\Box\phi-\frac{1}{6}\,R\phi=\frac{\lambda}{3}\,\phi^3.\label{vac-kg-eq}\eea This means that even in the vacuum -- supposedly scale invariant -- gravity (\ref{deser-action}), if probe the gravitational laws by means of test particles, the difference between gauges will be revealed.
The conclusion is trivial: since under the Weyl gauge transformations (\ref{scale-t}) the equations of motion of matter -- represented either by the conservation of stress-energy for matter fluxes or by the geodesic equations for point-like particles -- are indeed modified, the theory (\ref{deser-action}) (the same for (\ref{bars-action})) is not actually Weyl-invariant. Only the laws of gravity are unchanged by (\ref{scale-t}). As we shall see in the next section, this is a consequence of adopting spacetime manifolds whose geometrical structure is dictated by Riemann geometry, as the arena where to play the gravitational laws governed by (\ref{deser-action})/(\ref{bars-action}). The lesson to be learned is that there will be problems with a theory which pretends to be Weyl-invariant only because the action -- and the derived field equations -- is invariant under (\ref{scale-t}), but which is sustained by spacetimes whose geometrical structure does not share the gauge symmetry of the action.
\subsection{Gauge freedom}
The lack of Weyl gauge symmetry in the theory (\ref{deser-action}) is better illustrated by the following discussion. In the above vacuum field equations which are derived from (\ref{deser-action}), the KG equation (\ref{vac-kg-eq}) coincides with the trace of (\ref{vac-feq}). Hence, one degree of freedom is redundant:\footnote{We want to point out that even if consider radiation, in which case the Einstein equations (\ref{feq}) are satisfied, given that the radiation is traceless, the KG equation (\ref{kg-eq}) is also redundant since it coincides with the trace of the Einstein equations. This means that in this case a redundant degree of freedom to make Weyl gauge transformations (\ref{scale-t}) also arises.} there are -- in principle -- 10 equations to solve for 11 unknowns $g_{\mu\nu}$, $\phi$. The redundant degree of freedom is due to scale invariance of the action (\ref{deser-action}). What this means is that the vacuum field equations are not enough to determine the dynamics of the gravitational fields $(g_{\mu\nu},\phi)$. A feasible way out is to choose one specific gauge, say the GR gauge, or, as it is also called, the Einstein gauge (E-gauge) where $\phi_\text{E}=\sqrt{3/4\pi}\,G^{-1/2}$, with $G$- the Newton's constant. Then $g^\text{E}_{\mu\nu}$ is the metric which solves the Einstein-de Sitter equation $$G_{\mu\nu}=\frac{3\lambda}{8\pi G}\,g_{\mu\nu}.$$ The whole class of conformal spaces which solve (\ref{vac-feq}) can be then generated:
\bea {\cal C}_\text{E}=\{(g^{(i)}_{\mu\nu},\phi_{(i)}):g^{(i)}_{\mu\nu}=\phi_{(i)}^{-2}g^\text{E}_{\mu\nu},\;\phi_{(i)}=f_i(x)\},\label{c-class}\eea where the $f_i$ are in the infinite countable set of the (non-vanishing) real-valued smooth functions. In other words, one has a whole (infinite) class of pairs $(g^{(i)}_{\mu\nu},\phi_{(i)})$ -- of spacetimes $({\cal M},g^{(i)}_{\mu\nu})$ correspondingly -- which satisfy the vacuum field equations (\ref{vac-feq}). The different spacetimes in the above class were equivalent geometrical representations of a given gravitational phenomenon if it were not for the 2nd term in the RHS of (\ref{non-geod}) which can be understood as an additional ``5-force'' $\propto-\Omega^{-1}\der^\mu\Omega$ of non-gravitational origin. Actually, given that in the E-gauge the test point-particles follow the Riemannian geodesics (\ref{riem-geod}), i. e., there is not additional 5-force in the E-gauge, in any of the conformal gauges in ${\cal C}_\text{E}$ given by Eq. (\ref{c-class}), the 5-force $\propto\phi_{(i)}^{-1}\der^\mu\phi_{(i)}$ arises which has different strength in the different conformal gauges. Hence, in principle, by doing experimentation with (time-like) test point-particles one is able to rule-out those gauges which do not meet the tight experimental bounds on 5-force \cite{will}. What this means is that (\ref{deser-action}) -- the same for (\ref{bars-action}) -- represents, in fact, a whole class of different theories related by Weyl gauge transformations (\ref{scale-t}). The same conclusion is obviously true for the theory (\ref{tot-action}) where matter degrees of freedom are considered.
\subsection{Spacetime singularities}\label{sing-sub}
The additional non-gravitational force in (\ref{non-geod}) is the responsible for the avoidance -- do not confound with removal -- of certain spacetime singularities which might be present in the original spacetime but which are not met in the conformal one: incomplete time-like geodesics in the original spacetime can be mapped into complete (non)geodesics in the conformal space thanks to the 5-force $\propto-\Omega^{-1}\der^\mu\Omega$, which deviates the motion of a time-like test particle from being geodesic. The role of the conformal transformations in this case is to send the singularity to one of the ends of the (complete) time-like non-geodesic curve at infinity. I. e., the singularity is still there but it takes an infinite proper (conformal) time along one such time-like non-geodesic curve to reach to it \cite{quiros-2000, kaloper}. Notice that we repeat ``time-like'' every time to emphasize that the given spacetime singularity may be hidden from time-like test observers but not from photons (in general from massless particles). Actually, the geodesic equations for massless particles are not affected by the conformal transformation of the metric \cite{wald}, so that the radiation -- this includes the gravitational waves -- always sees the singularity (see \cite{kaloper} for a related discussion).
That the singularity which exists in one gauge is not removed in the conformal gauges can be seen if take a look at the Weyl gauge curvature invariants \cite{carrasco-kallosh}. One may argue that, working as we do with Riemannian spacetimes, it could be enough to explore the Riemannian invariants
\bea &&I_2\equiv R,\;I_4\equiv R_{\mu\nu}R^{\mu\nu},\;I'_4\equiv R_{\mu\nu\kappa\lambda}R^{\mu\nu\kappa\lambda},\nonumber\\
&&\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;I''_4\equiv C^2:=C_{\mu\nu\kappa\lambda}C^{\mu\nu\kappa\lambda},\label{riem-inv}\eea where $C_{\mu\nu\kappa\lambda}$- the Weyl tensor, etc, but these are not Weyl gauge invariant quantities. Instead, the following curvature invariants
\bea I^\phi_2\equiv\phi^{-4}\left[\phi^2R+6(\der\phi)^2\right],\;I''^\phi_4\equiv\phi^{-4}C_{\mu\nu\kappa\lambda}C^{\mu\nu\kappa\lambda},\label{weyl-g-inv}\eea are actual Weyl gauge invariants. To construct curvature invariants related with the Ricci and the Riemann-Christoffel tensors is by far a more complex task \cite{kallosh-linde}.
By construction the gauge-free curvature scalars $I^\phi_2$ and $I''^\phi_4$ in (\ref{weyl-g-inv}) are invariant scalars both under general coordinate transformations and under Weyl gauge transformations (\ref{scale-t}), in other words, given that in the E-gauge, $$I^\phi_2=\phi^{-2}_\text{E}R(g_\text{E})=\frac{4\pi}{3}\,GR_\text{E},\;I''^\phi_4=\frac{16\pi^2}{9}\,G^2C^2_\text{E},$$ then
\bea &&I^\phi_2=\frac{4\pi}{3}\,GR_\text{E}=\phi^{-4}\left[\phi^2R+6(\der\phi)^2\right],\nonumber\\
&&I''^\phi_4=\frac{16\pi^2}{9}\,G^2C^2_\text{E}=\phi^{-4}C^2.\nonumber\eea This means that if there is a curvature singularity at some spacetime point where $R_\text{E}$ and/or $C^2_\text{E}$ blow up in the E-gauge, this curvature singularity will be present in any other gauge. In Ref. \cite{carrasco-kallosh}, for instance, the Weyl gauge invariant $I''^\phi_4$ was used to show that any existing singularity in the E-gauge of the Weyl-invariant theory of \cite{bars, bars', bars''} will be also there in any of its Weyl conformal gauges (see a differing viewpoint in Ref. \cite{bars-ult}). As a matter of fact, given that the Weyl tensor governs the propagation of gravitational radiation through free space, this is a convenient quantity to judge about spacetime singularities in vacuum gravity. In a similar fashion, several years ago it was demonstrated in \cite{kaloper} that even if time-like free-falling observers in a conformal frame avoid a given existing singularity in the original frame, the null geodesics can not avoid hitting the singularity in a finite time since these are not modified by the conformal transformations. In other words, a given existing singularity can be hidden from time-like observers in the conformal frame, but it can not be hidden from massless particles (in \cite{kaloper} this was demonstrated for gravitational radiation).
Other gauge-free quantities may be constructed which are related with powers of scalar densities, instead of being just scalars. Take, for instance \cite{bars, bars', bars'', bars-ult, kallosh-linde, carrasco-kallosh}: $${\cal J}_0\equiv |g|^{1/4}\frac{\phi^2-\sigma^2}{6}.$$ In this case a power of the square of the metric determinant $\sqrt{|g|}$ is implied. This is a scalar density of weight $+1$: $\sqrt{|g'|}=J\sqrt{|g|}$, where $J=\det[\der x^\lambda/\der x'^\kappa],$ is the Jacobian determinant. It is included as part of the measure in the action integral in combination with the volume element: $d^4x\sqrt{|g|}$, to allow for general coordinate invariance. Since under general coordinate transformations the scalar densities are in general transformed, the utility of gauge-invariant quantities like ${\cal J}_0$ in the study of the spacetime singularities is unclear. To illustrate this let us assume the FRW metric with flat spatial sections. Given the peculiar properties of the cosmological metric, it is $dt\sqrt{|g|}$ which transforms like $\sqrt{|g|}$ for a general metric, under a conformal transformation $g_{\mu\nu}\rightarrow\Omega^{-2}g_{\mu\nu}$. Actually, let us write the FRW (flat) metric in the Einstein frame and in any other conformal frame $g^E_{\mu\nu}=\Omega^2g_{\mu\nu}$: $$ds_\text{E}^2=-dt_\text{E}^2+a_\text{E}^2(t_\text{E})d{\bf x}^2,\;ds^2=-dt^2+a^2(t)d{\bf x}^2,$$ where $$dt=\Omega^{-1}dt_\text{E},\;a(t)=\Omega^{-1}a_\text{E}(t_\text{E}),$$ then $$\sqrt{|g|}=\Omega^{-3}\sqrt{|g_E|},\;dt\sqrt{|g|}=\Omega^{-4}dt_E\sqrt{|g_E|}.$$
Misunderstanding of this fact may lead to the following situation. Let us assume a single conformally coupled scalar field like in (\ref{tot-action}). For the flat FRW metric we have that ${\cal J}_0=a^{3/2}\phi^2/6.$ It is known that the action (\ref{tot-action}) is mapped into the GR Einstein-Hilbert action if choose $\Omega=\phi/\phi_E$, where, for simplicity, we assume that the constant $\phi_E=1$. Hence $dt=\phi^{-1}dt_\text{E}$, $a(t)=\phi^{-1}a_\text{E}(t_\text{E})$. Since $\sqrt{|g_E|}=a_E^3$, from the definition above we have that (recall that $\phi_E=1$) $${\cal J}^E_0=a_E^{3/2}/6,\;{\cal J}_0=a^{3/2}\phi^2/6.$$ If assume that ${\cal J}_0$ is an actual invariant as it is done in \cite{bars, bars', bars'', bars-ult}, then $${\cal J}^E_0={\cal J}_0\;\Rightarrow\;a=\phi^{-4/3}a_E,$$ which is incorrect. It is obvious that ${\cal J}_0$ is a Weyl-gauge invariant, but, as said, it is not an invariant under general coordinate transformations, and in the above equations, due to the peculiar properties of the cosmological metric, we have taken into account that $dt=\phi^{-1}dt_E$, which is a coordinate transformation. Besides, if assume the theory (\ref{tot-action}) with $\lambda=0$ and ${\cal L}_\text{mat}={\cal L}_\text{rad}$ -- the Lagrangian for radiation to govern the cosmological dynamics of flat FRW spacetimes, then, since $3H^2_E=\rho_\text{rad}\propto a^{-4}_E$ $\Rightarrow\;a_E\propto t_E^{1/2}$, there is a real cosmological singularity at $t_E=0$, where $\rho_\text{rad}\propto t^{-2}_E$ blows up, even if $${\cal J}_0=a^{3/2}_E\phi^2_E=a^{3/2}_E\propto t^{3/4}_E,$$ is finite at $t_E=0$.
Likewise the quantity ${\cal J}_0$ one may invent many other such gauge-invariant combinations like, for instance,
\bea &&{\cal J}_1\equiv\sqrt{|g|}\left[\frac{\left(\phi^2-\sigma^2\right)}{12}\,R+\frac{1}{2}(\der\phi)^2-\frac{1}{2}(\der\sigma)^2\right],\nonumber\\
&&{\cal J}_2\equiv\sqrt{|g|}C_{\mu\nu\kappa\lambda}C^{\mu\nu\kappa\lambda},\nonumber\eea etc, but, what is the point of this? We have genuine gauge-free curvature invariants (scalars) like $I^\phi_2$ and $I''^\phi_4$ above, which may decide whether a given singularity is a true curvature singularity or just a coordinate or gauge artifact (see the discussion in section \ref{sing}).
In the next sections we shall explore a feasible slight modification of the Riemann spacetime structure in the search for a different geometrical arena where to play the scale-invariant laws of gravity. It will be shown, in particular, that the search for gauge-independent curvature invariants related with the curvature scalar, and with the Ricci, Riemann and Weyl tensors is by far a much more simple task.
\section{Weyl-integrable geometry}\label{wing}
In the former sections we have assumed -- as it is done usually in the bibliography on scale invariance (see, for instance \cite{bars, bars', bars'', kallosh-linde, carrasco-kallosh, bars-ult, padilla, percacci, prester, shaposhnikov}) -- that the affine properties of the underlying spacetimes are governed by Riemann geometry. However, in Riemannian spacetimes, given the transformation properties of the Riemannian affine connection (properly the Christoffel symbols of the metric) depicted by (\ref{1-vp}), the affine properties are not preserved by the Weyl gauge transformations (\ref{scale-t}).
Let us look for feasible modifications of Riemann geometry which could accommodate Weyl gauge symmetry. The first (and simplest) such modification that comes to one's mind is Weyl geometry \cite{weyl, perlick, novello, scholz-h, bib-weyl, books, chinos, chinos-1}. Weyl's geometric theory is a minimal generalization of Riemann geometry to include point dependent length of vectors during parallel transport, in addition to the point dependent property of vectors directions.\footnote{In Ref. \cite{chinos} the conformal invariance in Einstein-Cartan-Weyl spaces is investigated, while in \cite{chinos-1} Weyl invariance is extended to anisotropic spacetime version.} It is assumed that the length of a given vector ${\bf l}$ ($l\equiv\sqrt{g_{\mu\nu} l^\mu l^\nu}$) which is submitted to parallel transport varies from point to point in spacetime according to: $dl=l w_\mu dx^\mu/2$, where $w_\mu$ is the Weyl gauge boson. Here we shall concentrate in a special branch of Weyl geometry dubbed as Weyl-integrable geometry or WING for short. WING is obtained from Weyl geometry by replacing $w_\mu\rightarrow\der_\mu w$, where $w$ is the Weyl gauge scalar.\footnote{In this case, since $\oint dx^\mu \der_\mu w/2=0$, then the lengths of vectors, although point-dependent, are integrable. The units of measure in WING manifolds are point-dependent quantities unlike in Riemannian spaces. For a concise exposition of the fundamentals of Weyl-integrable geometry we submit the reader to the classical texts \cite{books}.}
Here we shall consider the action (\ref{deser-action}), and we will postulate that the spacetimes which solve the derived field equations have WING affine structure in place of the Riemannian structure assumed in the former sections. In order to make the theory (\ref{deser-action}) compatible with the former postulate we shall assume that the dilaton $\phi$ is the Weyl gauge scalar -- the precise correspondence with the well-known Weyl boson is actually $w=\ln\phi^2$ -- so that it is lifted to the category of a geometric field in addition to the metric field itself $g_{\mu\nu}$. The corresponding WING affine connection of the manifold $\Gamma^\alpha_{\beta\mu}$ is defined as
\bea \Gamma^\alpha_{\beta\mu}\equiv\{^\alpha_{\beta\mu}\}+\frac{1}{\phi}\left(\delta^\alpha_\beta\der_\mu\phi+\delta^\alpha_\mu\der_\beta\phi-g_{\beta\mu}\der^\alpha\phi\right).\label{wig-aff-c}\eea The resulting Weyl-invariant action -- which is coupled to WING background spacetimes -- reads
\bea S^{(w)}=\frac{1}{12}\int d^4x\sqrt{|g|}\left[\phi^2R^{(w)}+\lambda\phi^4\right],\label{deser-mod}\eea where the label ``$(w)$'' refers to Weyl-integrable quantities, which are defined with respect to the affine connection (\ref{wig-aff-c}). Notice that the kinetic term for the Weyl gauge boson (former dilaton) $\phi$ is already included in the definition of the WING-curvature scalar (up to a divergence) $$R^{(w)}=R+6\frac{(\der\phi)^2}{\phi^2},$$ where in the RHS of this equation the given quantities and operators coincide with their Riemannian definitions in terms of the Christoffel symbols of the metric.
One may as well consider adding a non-geometric, minimally coupled scalar field $\sigma$, which under (\ref{scale-t}) transforms like $\sigma\rightarrow\Omega\sigma$. The resulting Weyl invariant action can be written as follows (here we omit writing the quartic potential term $\propto\phi^4$):
\bea S^{(w)}=\int d^4x\sqrt{|g|}\left\{\frac{1}{12}\,\phi^2R^{(w)}-\frac{1}{2}(D\sigma)^2\right\},\label{deser-mod'}\eea where, as before, the different geometric quantities and operators labeled with the ``$(w)$'' refer to WING objects, and the standard derivative of the non-geometric field minimally coupled to gravity $\sigma$, have been replaced by the gauge-covariant derivative $$\der_\mu\sigma\rightarrow D_\mu\sigma\equiv\left(\der_\mu-\frac{\der_\mu\phi}{\phi}\right)\sigma.$$ Following this procedure one might add any number of minimally coupled scalars.
This apparently slight modification of (\ref{deser-action}) designed to make that theory compatible with WING backgrounds, results in that scale invariance is an actual symmetry of the laws of gravity. First of all notice that the torsionless affine connection of the WING space (\ref{wig-aff-c}), as well as the non-metricity condition
\bea \nabla^{(w)}_\mu g_{\kappa\lambda}=-2\frac{\der_\mu\phi}{\phi}\,g_{\kappa\lambda}\;\Rightarrow\;\nabla_{(w)}^\kappa g_{\kappa\lambda}=-2\frac{\der^\kappa\phi}{\phi}\,g_{\kappa\lambda},\label{wing-law}\eea which is the fundamental geometric law of WING spaces, both are unchanged by the scale transformations (\ref{scale-t}). As a consequence, the WING-Ricci tensor $R^{(w)}_{\mu\nu}$ and the corresponding WING-Einstein's tensor $G^{(w)}_{\mu\nu}\equiv R^{(w)}_{\mu\nu}-g_{\mu\nu} R^{(w)}/2$, as well as the covariant derivative operator $\nabla^{(w)}_\mu$, etc, are scale-invariant objects.
In this Weyl-invariant modification of (\ref{deser-action}) which is grounded in WING backgrounds, not only the field equations derived from (\ref{deser-mod})/(\ref{deser-mod'}) -- see below -- but also the WING-geodesics $$\frac{d^2x^\alpha}{ds^2}+\Gamma^\alpha_{\mu\nu}\frac{dx^\mu}{ds}\frac{dx^\nu}{ds}=\frac{\der_\mu\phi}{\phi}\frac{dx^\mu}{ds}\frac{dx^\alpha}{ds},$$ or after a convenient re-parametrization $d\sigma=\phi\,ds$,
\bea \frac{d^2x^\alpha}{d\sigma^2}+\Gamma^\alpha_{\mu\nu}\frac{dx^\mu}{d\sigma}\frac{dx^\nu}{d\sigma}=0,\label{mod-geod}\eea and the WING continuity equation
\bea \nabla_{(w)}^\kappa T^\text{mat}_{\kappa\mu}=2\frac{\der^\kappa\phi}{\phi}\,T^\text{mat}_{\kappa\mu},\label{mod-cons-eq}\eea all are invariant under the Weyl gauge transformations (\ref{scale-t}). In equation (\ref{mod-cons-eq}) -- compare with the non-metricity condition in the form of the right-hand side Eq. (\ref{wing-law}) -- the term in the RHS is not actually a source term, instead it expresses the fact that in WING spaces the units of measure of energy and stresses are point-dependent quantities, as well as the length of any other vector. As a matter of fact the above conservation equation looks similar to that in GR if redefine the stress-energy tensor of matter $T^{w,\text{mat}}_{\mu\nu}\equiv\phi^{-2}T^\text{mat}_{\mu\nu}$,
\bea \nabla_{(w)}^\kappa T^{w,\text{mat}}_{\kappa\mu}=0.\label{wig-cont-eq}\eea This is the stress-energy tensor which has the Weyl-invariant physical meaning since, under (\ref{scale-t}), it is unchanged like the WING-Einstein's tensor $G^{(w)}_{\mu\nu}$.
\subsection{Field equations}
Let us to explore an specific example. Consider the following action \cite{quiros-2014}:
\bea &&S^{(w)}_\text{example}=\int d^4x\sqrt{|g|}\left\{\frac{1}{12}\,\phi^2R^{(w)}-\right.\nonumber\\
&&\left.\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\frac{1}{2}|D\sigma|^2-\frac{\lambda}{4}\left(|\sigma|^2-v^2_0\phi^2\right)^2\right\},\label{ex-action}\eea where $\sigma$ can be the Higgs field in the unitary gauge, in which case $v_0$ is the mass parameter of the standard model of particles, and $\lambda$ is a coupling constant. The above action is invariant under (\ref{scale-t}), plus the Higgs field transformation $\sigma\rightarrow\Omega\,\sigma$. In order to derive the field equations it is convenient to introduce a gauge-independent scalar field variable $\chi=\phi^{-1}\sigma$. After this the action (\ref{ex-action}) is rewritten as
\bea S^{(w)}_\text{example}=\int d^4x\sqrt{|g|}\left\{\frac{1}{12}\,\phi^2R^{(w)}+\phi^2{\cal L}_\text{Higgs}\right\},\label{chi-action}\eea where the Higgs field Lagrangian is given by
\bea {\cal L}_\text{Higgs}=-\frac{1}{2}(\der\chi)^2-\frac{\lambda\phi^2}{4}(\chi^2-v^2_0)^2.\label{higgs-lag}\eea
Written in the form (\ref{chi-action}) the action of the theory is manifestly Weyl-invariant, where we recall that under the scale transformations (\ref{scale-t}) the field $\chi$ is unchanged, i. e., it is already a Weyl gauge invariant scalar. Notice also that, unlike in (\ref{deser-action})/(\ref{bars-action}) the minimal coupling of the Higgs field to gravity does not spoil the scale invariance of the action. The following manifestly scale-invariant field equations are easily derived from (\ref{chi-action}):
\bea &&G_{\mu\nu}^{(w)}=\frac{1}{M^2_\text{pl}}\,T^{(w,\chi)}_{\mu\nu}\;\Rightarrow\nonumber\\
&&G_{\mu\nu}-\frac{2}{\phi}(\nabla_\mu\der_\nu\phi-g_{\mu\nu}\Box\phi)+\nonumber\\
&&\;\;\;\;\;\;\;4\frac{\der_\mu\phi}{\phi}\frac{\der_\nu\phi}{\phi}-g_{\mu\nu}\frac{(\der\phi)^2}{\phi^2}=\frac{1}{M^2_\text{pl}}\,T^{(w,\chi)}_{\mu\nu},\label{e-feq}\\
&&\nabla^\kappa_{(w)}T^{(w,\chi)}_{\kappa\mu}=0\;\Rightarrow\nonumber\\
&&\Box\chi+2\frac{(\der\phi\cdot\der\chi)}{\phi}=\lambda\phi^2\left(\chi^2-v^2_0\right)\chi,\label{chi-kg-eq}\eea where in the second lines (explicit form) of the above equations the geometric objects and operators are the usual Riemannian quantities, $M^2_\text{pl}=1/6$ in the units chosen in this paper, and
\bea &&T^{(w,\chi)}_{\mu\nu}=-\frac{2}{\sqrt{|g|}}\frac{\der(\sqrt{|g|}\,{\cal L}_\text{Higgs})}{\der g^{\mu\nu}}\nonumber\\
&&\;\;\;\;\;\;\;\;\;\;\;=\der_\mu\chi\der_\nu\chi-\frac{1}{2}g_{\mu\nu}(\der\chi)^2\nonumber\\
&&\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;-\frac{\lambda\phi^2}{4}\,g_{\mu\nu}\left(\chi^2-v^2_0\right)^2,\label{chi-set}\eea is the already Weyl-invariant stress-energy tensor of the Higgs field. If derive the KG equation for the Weyl gauge scalar $\phi$ (former dilaton) from the action (\ref{chi-action}), one obtains $$\Box\phi-\frac{1}{6}R\phi=\phi\,T^{(w,\chi)}=-\phi\left[(\der\chi)^2+\lambda\phi^2\left(\chi^2-v^2_0\right)^2\right],$$ which exactly coincides with the trace of the Einstein's equations (\ref{e-feq}) independent of the type of matter under consideration. Recall that in the standard situation in the references \cite{bars, bars', bars'', kallosh-linde, carrasco-kallosh, bars-ult, padilla, percacci, prester, shaposhnikov} where Riemannian backgrounds are adopted, only traceless matter can be consistently added to the action (\ref{deser-action}) (same for (\ref{bars-action})).
Summarizing: the adoption of Weyl-integrable geometry as the theory which governs the affine properties of the spacetime background allows to construct a fully Weyl-invariant theory even if matter degrees of freedom other than radiation (and non-minimally coupled scalars) are considered. The subtle idea was to lift the dilaton $\phi$ to the category of a geometric field, i. e., to adopt it as the Weyl gauge boson $w=\ln\phi^2$ of WING, i. e., the one which takes part in the definition of the affine connection (\ref{wig-aff-c}) and consequently in the definition of the curvature of spacetime.
\subsection{Gauge freedom}
As already mentioned, the KG equation for the Weyl gauge boson $\phi$ is not an independent equation, but it coincides with the trace of the WING-Einstein equations (\ref{e-feq}). This is a direct consequence of scale invariance which means that there is not just a single Weyl-integrable space $({\cal M},g_{\mu\nu},\phi)$ -- properly a gauge -- which is solution of (\ref{e-feq}), (\ref{chi-kg-eq}), but a whole equivalence class of them:
\bea &&{\cal C}=\{(g^{(a)}_{\mu\nu},\phi_{(a)}):a=1,2,...,i,...,k,...,\infty;\nonumber\\
&&\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;g^{(i)}_{\mu\nu}=\Omega^2_{ik}g^{(k)}_{\mu\nu},\;\phi_{(i)}=\Omega^{-1}_{ik}\phi_{(k)}\},\nonumber\eea i. e., any two pairs in ${\cal C}$: $(\bar g_{\mu\nu},\bar\phi)$ and $(g_{\mu\nu},\phi)$, are related by a scale transformation $\bar g_{\mu\nu}=\Omega^2g_{\mu\nu}$, $\bar\phi=\Omega^{-1}\phi$. As it was illustrated in the cosmological context in \cite{quiros-2014}, the most one can get from the field equations is a functional relationship among the gravitational potentials $g_{\mu\nu}$ and $\phi$. This relationship is independent of the matter content: for a given matter source one have an infinity of possibilities $(g^{(a)}_{\mu\nu},\phi_{(a)})$, where $a=1,2...,\infty$ and the $\phi_{(a)}$ belong in the space of continuous real-valued functions. Each possible gauge $(g^{(i)}_{\mu\nu},\phi_{(i)})\in{\cal C}$, represents a potential geometric description of the laws of gravity (and of particle's physics). From the cosmological standpoint, for instance, to have an infinity of feasible -- fully equivalent -- geometrical descriptions amounts to have an infinity of possible patterns of cosmological evolution which satisfy the same cosmological field equations. For a detailed discussion of this issue and possible connections with the multiverse picture \cite{multiverse} see \cite{quiros-2014}.
The simplest gauge one may choose is the one where $\phi=\phi_\text{E}=$ const. This trivial E-gauge corresponds to general relativity since, after the above choice, the action (\ref{ex-action}) transforms into the Einstein-Hilbert (EH) action minimally coupled to the standard model of particles with no new physics beyond the standard model at low energies. At the same time WING spaces are transformed into pseudo-Riemannian manifolds.
It is a very simple exercise to show that the spacetimes which solve the GR field equations belong in the conformal class ${\cal C}$. Actually, let us consider the WING-EH action (\ref{deser-mod}) $$S^{(w)}_\text{EH}=\frac{1}{12}\int d^4x\sqrt{|g|}\left[\phi^2R^{(w)}+\lambda\phi^4\right].$$ Under a Weyl rescaling (\ref{scale-t}) with $\Omega^2=\phi^2$, the WING affine connection (\ref{wig-aff-c}) is transformed into the Christoffel symbols of the conformal metric: $$\Gamma^\mu_{\alpha\beta}\rightarrow\{^\mu_{\alpha\beta}\}\;\Rightarrow\;R^{(w)}_{\mu\nu}\rightarrow R_{\mu\nu},\;\text{etc.}$$ then $$S^{(w)}_\text{EH}\rightarrow S_\text{EH}=\frac{1}{12}\int d^4x\sqrt{|g|}\left[R+\lambda\right].$$ Through an inverse transformation (\ref{scale-t}) with $\Omega^2=\phi^{-2}$, each GR solution generates an infinite set of spacetimes back in ${\cal C}$. To see this let us start with a given known GR solution $({\cal M},g^\text{E}_{\mu\nu})$, and perform the conformal transformation of the GR metric: $$g^\text{E}_{\mu\nu}\rightarrow\Omega^2g^\text{E}_{\mu\nu},\;\Omega^2=\phi^{-2}.$$ Under the above transformation, the Riemannian quantities and operators are mapped back into WING objects: $$\{^\mu_{\alpha\beta}\}\rightarrow\Gamma^\mu_{\alpha\beta}\;\Rightarrow\;R_{\mu\nu}\rightarrow R^{(w)}_{\mu\nu},\;\text{etc.}$$
It is not difficult to realize that there exists an infinite countable set of smooth real-valued functions $\phi_{(a)}=f_a(x)$, such that every pair $$(g^{(a)}_{\mu\nu},\phi_{(a)}),\;\text{with}\;g^{(a)}_{\mu\nu}=\phi_{(a)}^{-2}g^\text{E}_{\mu\nu},\;a=1,2,...\infty,$$ belongs in ${\cal C}$. This means that each GR spacetime solution $({\cal M},g^\text{E}_{\mu\nu})$, together with its infinite set of equivalent conformal representations $$\{({\cal M},g^{(a)}_{\mu\nu},\phi_{(a)}):\,g^{(a)}_{\mu\nu}=\phi_{(a)}^{-2}g^\text{E}_{\mu\nu},\;a=1,2,...\infty\},$$ belong in the equivalence class ${\cal C}$.
\section{Gauge-free curvature invariants and spacetime singularities}\label{sing}
There are issues which require of gauge-independent geometric invariants to reach to gauge-independent conclusions. An outstanding example is the singularity issue. There has been a debate about the possibility that certain spacetime singularities in scalar-tensor theories can be avoided in their conformal formulations \cite{quiros-2000, faraoni, kaloper}. Unlike scalar-tensor theories which are not scale invariant, the theory (\ref{ex-action}) is Weyl gauge-invariant so that the discussion of this issue is clearly different.
In order to discuss on the occurrence of spacetime singularities in the above setup one is obliged to resort to the geometric invariants like
\bea &&I^\phi_2\equiv\phi^{-2}R^{(w)},\;I^\phi_4=\phi^{-4}R^{(w)}_{\mu\nu}R^{\mu\nu}_{(w)},\nonumber\\
&&I'^\phi_4\equiv\phi^{-4}R^{(w)}_{\mu\nu\kappa\lambda}R^{\mu\nu\kappa\lambda}_{(w)},\label{wing-inv}\eea etc, which are not transformed by the scale transformations (\ref{scale-t}) and, hence, are the ones which carry gauge-independent physical meaning.
As it was demonstrated in the former section, general relativity is a particular gauge of the theory (\ref{ex-action}) when the Weyl scalar $\phi=\phi_\text{E}$ is a constant where, without lost of generality, we may set $\phi_\text{E}=1$. This entails that the following equalities involving the gauge-independent curvature invariants (\ref{wing-inv}) are satisfied:
\bea &&I^\phi_2=R_\text{E}=\phi^{-2}R^{(w)},\nonumber\\
&&I^\phi_4=R^\text{E}_{\mu\nu} R_\text{E}^{\mu\nu}=\phi^{-4}R^{(w)}_{\mu\nu}R^{\mu\nu}_{(w)},\nonumber\\
&&I'^\phi_4=R^\text{E}_{\mu\nu\kappa\lambda} R_\text{E}^{\mu\nu\kappa\lambda}=\phi^{-4}R^{(w)}_{\mu\nu\kappa\lambda}R^{\mu\nu\kappa\lambda}_{(w)},\label{sing-inv}\eea where the quantities with an ``E'' denote Riemannian objects defined in terms of the Christoffel symbols of the metric $g^\text{E}_{\mu\nu}$ in the E-gauge. Notice that equalities like the ones in (\ref{sing-inv}) arise only in gauge invariant theories where the gauge-independent curvature invariants (\ref{wing-inv}) make sense. These are not necessarily satisfied when dealing with standard scalar-tensor theories like the BD theory \cite{bd}.
Let us assume the following hypothetical situation: the spacetime in the E-gauge $({\cal M},g^\text{E}_{\mu\nu})$ has a curvature singularity at some point $x_P$ in the manifold, such that the GR invariant $$I'^\phi_4=R^\text{E}_{\mu\nu\kappa\lambda} R_\text{E}^{\mu\nu\kappa\lambda}=\alpha(x),$$ blows up at $x_P$, i. e.,
\bea \lim_{x\rightarrow x_P}\alpha(x)\rightarrow\infty.\label{lim}\eea Let us further assume that $$\lim_{x\rightarrow x_P}\phi^4\rightarrow 0,$$ where the latter limit is approached in such a way that the WING-curvature invariant $$R^{(w)}_{\mu\nu\kappa\lambda}R^{\mu\nu\kappa\lambda}_{(w)}=\alpha\phi^4,$$ remains finite at $x_P$. This means that, given that the conditions of the hypothetical situation described above are fulfilled, the curvature singularity in the E-gauge is not felt by an observer living in the conformal WING-world. This does not mean that the singularity has been erased by the Weyl rescaling
\bea g_{\mu\nu}=\Omega^{-2}g^\text{E}_{\mu\nu},\;\phi=\Omega\phi_\text{E}.\label{scale-t'}\eea As a matter of fact the physically meaningful gauge-independent curvature invariant $I'^\phi_4=\alpha$ blows up at $x_P$, meaning that the singularity is still there. To understand what actually happens when the singularity is approached one have to recall that in WING spacetimes the units of measure are point-dependent. As the singularity is approached and the given gauge-independent curvature invariant, say $I'^\phi_4$, grows up without bound, the corresponding units of measure increase in a similar fashion so that the increase in the magnitude of the invariants is conveniently balanced.
\subsection{Big-gang singularity}
For a more specific qualitative analysis it is convenient to study a concrete example. Here we shall consider as an specific example the GR cosmological singularity usually associated with the big-bang. We have $$ds_\text{E}^2=-dt_\text{E}^2+a_\text{E}^2(t_\text{E})d{\bf x}^2,\;ds^2=-dt^2+a^2(t)d{\bf x}^2,$$ where the left-hand line-element refers to the (conformal) GR Friedmann-Robertson-Walker (FRW) spacetime, while the right-hand one refers to WIG-FRW spacetime. The following relationships arise:
\bea dt=\phi^{-1}dt_\text{E},\;a(t)=\phi^{-1}a_\text{E}(t_\text{E}),\;\rho^{(w)}=\phi^2\rho_\text{E},\label{rel}\eea where $\rho^{(w)}$ is the energy density of matter measured by a co-moving observer in the WING-world, and we assumed the conformal factor $\Omega^2=\phi^2$. Consider a singular GR solution $a_\text{E}\propto t_\text{E}^n$ ($n$ is an arbitrary positive constant), so that according to the GR-Friedmann constraint [$3H_\text{E}^2=\rho_\text{E}/M^2_\text{pl}$], we have $\rho_\text{E}\propto t_\text{E}^{-2}$. The initial cosmological singularity is at $t_\text{E}=0$ where the matter energy density $\rho_\text{E}$ blows up. It is a simple exercise to show that invariants $I^\phi_2$, $I^\phi_4$ and $I'^\phi_4$ go to infinity at the singularity. In particular $I^\phi_2\propto H_\text{E}^2\propto t_\text{E}^{-2}$ grows up without bound as $t_\text{E}\rightarrow 0$. Among the infinity of possibilities let us choose\footnote{Recall that due to the gauge freedom we are free to choose any $\phi$ we want.}
\bea \phi(t_\text{E})=\tanh(t_\text{E}).\label{vphi}\eea After the above convenient choice one gets that the WING curvature scalar $$R^{(w)}=\phi^2 I^\phi_2\propto\tanh^2(t_\text{E})/t^2_\text{E},$$ is always bounded: $0\leq t_\text{E}<\infty$ $\Rightarrow\,1\geq R^{(w)}>0$. The same is true for the energy density measured by a co-moving WING-observer $\rho^{(w)}=\phi^2\rho_\text{E}\propto R^{(w)}$. Hence a co-moving observer in the equivalent WING picture does not find singular behavior at all. As already stated, the explanation is simple: although the singularity is still there, as the co-moving observer approaches to it, any unbounded increment in any of the gauge-independent curvature invariants is balanced by a proportional increment in the corresponding units of measure in the WING-FRW spacetime.
Regarding the amount of cosmic time separating a given co-moving observer from the singularity we have to say that, while the interval of cosmic time from, say, the present moment of the cosmic history $t^\text{E}_0=t_0$ to the singularity in the past at $t_\text{E}=0$, is finite, in terms of the cosmic time measured by an observer in the WING-world we have that $$\Delta t=\int_\epsilon^{t_0}\frac{dt_\text{E}\cosh(t_\text{E})}{\sinh(t_\text{E})}=\ln\left(\frac{\sinh(t_0)}{\sinh(\epsilon)}\right),$$ where $\epsilon$ is a small number such that, as one approaches to the initial singularity, $\epsilon\rightarrow 0$. Hence, as $\epsilon\rightarrow 0$, $\Delta t\rightarrow\infty$. This means that to an observer in the WING-world the initial (big-bang) singularity is an infinite amount of cosmic time into the past. In consequence, for all practical purposes the corresponding WIG geodesics are complete into the past and the initial singularity -- although not erased -- is effectively avoided. Recall, however, that null-geodesics always meet the singularity in a finite time.
\section{Discussion and Conclusion}\label{conclu}
Although most theories of the fundamental interactions (including general relativity and string theory) assume that the geometric structure of spacetime is pseudo-Riemann, there are indications that a generalization of Riemann geometry -- notably Weyl integrable geometry -- might represent a better suited arena where to formulate the laws of gravity \cite{quiros-2013, quiros-2014, dicke, smolin, cheng, scholz}. As a matter of fact, it is surprising that Weyl geometry -- the natural arena for Weyl invariance -- is not usually adopted when dealing with scale invariance of the laws of gravity. Instead -- and regrettably -- any kind of discussion on the geometrical structure of spacetime is avoided when dealing with Weyl invariance. See, for instance, the incomplete list of references \cite{bars, bars', bars'', kallosh-linde, carrasco-kallosh, bars-ult, padilla, percacci, prester, shaposhnikov, deser} for a clear illustration of this statement.
Here we have learned that it is not enough to postulate a Weyl-invariant action like, for instance (see equations (\ref{deser-action}) and (\ref{bars-action})): $$S=\int d^4x\sqrt{|g|}\left[\frac{\phi^2}{12}\,R+\frac{1}{2}(\der\phi)^2\right],$$ or $$S=\int d^4x\sqrt{|g|}\left[\frac{\left(\phi^2-\sigma^2\right)}{12}\,R+\frac{1}{2}(\der\phi)^2-\frac{1}{2}(\der\sigma)^2\right],$$ which are invariants under the Weyl gauge transfromations $$g_{\mu\nu}\rightarrow\Omega^{-2}g_{\mu\nu},\;\phi\rightarrow\Omega\phi,\;\sigma\rightarrow\Omega\sigma.$$ Additionally one has to make a separate postulate on the geometrical laws which govern the affine structure of the background spacetime. If one postulates Riemann geometry -- as it is usually done --, since the Riemannian affine structure is modified by the Weyl gauge transformations, then scale invariance is condemned to fail. If, alternativelly, postulate that the affine properties of the space are dictated by Weyl-integrable geometry, then there is room for scale invariance to be an actual symmetry of the laws of gravity \cite{quiros-2013, quiros-2014}.
In section \ref{anomal} we have shown that, if associated with Riemannian spacetime backgrounds, the above actions (\ref{deser-action}) and (\ref{bars-action}) just coincide with Brans-Dicke theory with the anomalous value of the BD coupling constant $\omega=-3/2$. As it is well-known, in such a case only conformally coupled scalars and radiation do not explicitly destroy the Weyl invariance. We wrote ``explicitly'' because, as a matter of fact, even the vacuum theories depicted by these kinds of action are actually not conformally invariant theories. This is demonstrated by the fact that the motion equations for time-like test particles are not invariant under the Weyl gauge transformations. Of course, this is a consequence of the wrong choice of the laws of geometry: Riemannian geodesics are not invariant under the conformal transformations of the metric.
That misunderstanding of this fact may lead to wrong conclusions can be illustrated with the discussion on geodesic completeness within theories given by the action (\ref{bars-action}) (second equation at the beginning of this section) \cite{bars, bars', bars'', carrasco-kallosh, bars-ult}. In \cite{bars, bars', bars'', bars-ult} the authors rely on the study of time-like geodesics for particles with the mass $m$. They then made the argument that going to the limit of vanishing mass the completeness of null-geodesics can be incorporated in the discussion. Apparently, these authors forgot that their action is just BD gravity with BD coupling parameter $\omega=-3/2$, in which case only matter in the form radiation can be consistently considered, where by ``consistently'' we mean ``consistent with the derived field equations.'' A simple check of the field equations in their theory will clearly reveal that only radiation can be considered (see the related discussion in section \ref{anomal} of this paper). What is more disturbing: even if consider the vaccum theory, given that the background spacetimes are of Riemann affine structure -- which is actually modified by the conformal transformations -- the equations of motion of time-like test particles are transformed by the Weyl gauge transformations:
\bea &&\frac{d^2x^\mu}{ds^2}+\{^\mu_{\kappa\nu}\}\frac{dx^\kappa}{ds}\frac{dx^\nu}{ds}=0\;\rightarrow\nonumber\\
&&\frac{d^2x^\mu}{ds^2}+\{^\mu_{\kappa\nu}\}\frac{dx^\kappa}{ds}\frac{dx^\nu}{ds}=\frac{\der_\kappa\Omega}{\Omega}\frac{dx^\kappa}{ds}\frac{dx^\mu}{ds}-\frac{\der^\mu\Omega}{\Omega}.\nonumber\eea Only the null-geodesics are not transformed by the conformal transformations. But, this latter fact also contradicts the claimed geodesic completeness in \cite{bars, bars', bars'', bars-ult}: since the null-geodesics are not transformed by the Weyl gauge transformations, given that these are incomplete in the gauge where a given spacetime sigularity is present, these will be also incomplete in any of the conformally-related gauges. A similar argument signaling the incorrectness of the claim by the authors of \cite{bars, bars', bars''}, but using gauge-independent curvature invariants (in particular the one related with the Weyl invariant $C^2$), is found in \cite{carrasco-kallosh} (see the related discussion in section \ref{sing-sub} of this paper). The above controversy illustrates the confusion which may arise when exploring scale invariance if ignore the unavoidable discussion on the geometrical aspects of the given (metric) theory of gravity.
We have demonstrated that, perhaps, Weyl-integrable geometry -- being the result of lifting Riemann geometry to the category of a Weyl-invariant geometrical theory -- may be a better suited arena where to play the scale-invariant laws of gravity. Several consequences of such a fully Weyl-invariant theory of gravity and of particles for cosmology, as well as for particle's physics, are explored in \cite{quiros-2014}.
The author thanks David Stefanyszyn for helpful comments and the SNI of Mexico for support of this research.
|
\subsection{Notation}
For a graph $G$, $V(G)$ will be the vertex set of $G$ and $|G|=|V(G)|$. Where $W\subset V(G)$, $G[W]$ is the subgraph of $G$ induced on the vertices of $W$. The set of neighbours of a vertex $v$ is denoted by $N(v)$, and the neighbourhood of a vertex set $A\subset V(G)$ by $N(A)=(\cup_{v\in A}N(v))\setminus A$. Where multiple graphs are used, we use $N_G(v)$ for the neighbourhood of a vertex $v$ in the graph $G$. We use $N(A,B)$ to refer to the set of neighbours of $A$ in $B$, that is $N(A)\cap B$. We take $d_G(x, A)=|N_G(x,A)|$, and let
\[
d_G(A,B)=\sum_{x\in A}d_G(x,B).
\]
We say a path with $l$ vertices has \emph{length} $l-1$, and call a path $P$ an \emph{$x,y$-path} if the vertices $x$, $y$ have degree 1 in $P$. We call $x$, $y$ the \emph{ends} of $P$. When we remove a path from a graph we will remove all the edges of the path and delete any resulting isolated vertices.
Given two disjoint vertex sets $A$ and $B$, a \emph{$d$-matching from $A$ into $B$} is a collection of disjoint sets $\{X_a\subset N(a,B):a\in A\}$ so that, for each $a\in A$, $|X_a|=d$. As is well known, such matchings can be found by showing that Hall's generalised matching condition holds. For details on this, and other standard notation, see Bollob\'as~\cite{bollo1}. We use $\log$ for the natural logarithm, and in several places omit rounding signs when they are not crucial.
\subsection{Expanders and Almost Spanning Trees}
The main properties we will use for our embeddings will be various graph expansion properties. We use the same definition of expansion as Johannsen, Krivelevich and Samotij~\cite{JKS12}.
\begin{defn} Let $n\in \mathbb{N}$ and $d\in \mathbb{R}^+$. A graph $G$ is an \emph{$(n,d)$-expander} if $|V(G)|=n$ and $G$ satisfies the following two conditions.
\begin{enumerate}
\item $|N_G(X)|\geq d|X|$ for all $X\subset V(G)$ with $1\leq |X|<\lceil \frac{n}{2d}\rceil$.
\item $d_G(X,Y)>0$ for all disjoint $X,Y\subset V(G)$ with $|X|=|Y|=\lceil\frac{n}{2d}\rceil$.
\end{enumerate}
\end{defn}
Almost spanning trees can be found in expander graphs using a theorem of Haxell~\cite{PH01}, as shown by Balogh, Csaba, Pei and Samotij~\cite{BCPS10}. We will use the following formulation of this method, which differs from that used by Johannsen, Krivelevich and Samotij~\cite{JKS12} only in that a specific vertex of the tree is embedded to a specific vertex of the expander graph. Fortunately, this version follows identically by using the full statement of the theorem of Haxell.
\begin{theorem}\label{almost}
Let $n,\Delta\in \mathbb{N}$, let $d\in \mathbb{R}^+$ with $d\geq 2\Delta$, and let $G$ be an $(n,d)$-expander. Given any tree $T\in \mathcal{T}(n-4\Delta\lceil\frac{n}{2d}\rceil,\Delta)$ and vertices $v\in V(G)$ and $t\in V(T)$, we can find an embedding of $T$ in the graph $G$ with $t$ embedded on $v$.
\end{theorem}
For Theorem \ref{almostprob}, we will also require the following formulation, used by Balogh, Csaba, Pei and Samotij~\cite{BCPS10}.
\begin{theorem}\label{almost2}
Let $\Delta,m,M\in \mathbb{N}$. Let $H$ be a non-empty graph such that
\begin{enumerate}
\item if $X\subset V(H)$ and $0<|X|\leq m$, then $N_H(X)|\geq \Delta|X|+1$, and
\item if $m\leq |X|\leq 2m$, then $|N_H(X)|\geq 2\Delta|X|+M$.
\end{enumerate}
Then $H$ contains every tree in $\mathcal{T}(M,\Delta)$.
\end{theorem}
Expansion properties can also be used to construct Hamilton cycles, as shown by Hefetz, Krivelevich and Szab\'o~\cite{HKS09}.
\begin{theorem}\label{hamcycle}
Let $n,d\in \mathbb{N}$ satisfy that $n$ is sufficiently large and $12\leq d\leq e^{\sqrt[3]{\log n}}$. Then a graph $G$ on $n$ vertices is Hamiltonian if it satisfies the following conditions:
\begin{enumerate}
\item If $X\subset V(G)$ and $|X|\leq \frac{n\log\log n\log d}{d\log n\log\log\log n}$, then $|N(X)|\geq d|X|$,
\item If $X,Y\subset V(H)$ satisfy $|X|=|Y|\geq \frac{n\log \log n\log d}{4130\log n\log\log\log n}$ and are disjoint, then $d_G(X,Y)>0$.
\end{enumerate}
\end{theorem}
Specifically, we will use this theorem in the following form, which follows from Theorem \ref{hamcycle} by taking $d=\varepsilon\log n/\log\log n$.
\begin{corollary}\label{hamcyclesslightly}
Let $\varepsilon >0$. If $n$ is sufficiently large, and the graph $H$ is a $(n,\varepsilon\log n/\log\log n)$-expander, then $H$ is Hamiltonian.
\end{corollary}
\subsection{Finding paths}
Given many pairs of vertices in a graph and a simple large-set expansion property, we can find some path between one pair of the vertices using the following lemma.
\begin{lemma}[\cite{selflove1}] \label{connect} Let $m,n\in \mathbb{N}$ satisfy $m\leq n/800$, let $d=n/200m$ and let $n$ be sufficiently large. Let a graph $G$ with $n$ vertices have the property that any set $A\subset V(G)$ with $|A|=m$ satifies $|N(A)|\geq (1-1/64)n$. Suppose $G$ contains disjoint vertex sets $X$, $Y$ and $U$, with $X=\{x_1,\ldots,x_{2m}\}$, $Y=\{y_1,\ldots, y_{2m}\}$ and $|U|=\lceil n/8\rceil$. Suppose, in addition, we have integers $k_i$, $i\in[2m]$, satisfying $4\log n/\log d\leq k_i\leq n/40$. Then, for some $i$, there is an $x_i,y_i$-path of length $k_i$ whose internal vertices lie in $U$.
\end{lemma}
We say a set of subgraphs \emph{covers} a graph $G$ if every vertex in $G$ is contained in one of the subgraphs. To prove Theorem \ref{comblong}, we will need the following theorem.
\begin{theorem}[\cite{selflove1}]\label{pathcoverexpander} Let $n$ be sufficiently large and let $k\in \mathbb{N}$ satisfy $k\geq 10^3\log^3n$, $k|n$. Let a directed graph $G$ contain $n/k$ disjoint vertex pairs $(x_i,y_i)$ and let $W=V(G)\setminus (\cup_i\{x_i,y_i\})$. Suppose $G$ has the following two properties.
\begin{enumerate}
\item For any subset $A\subset V(G)$ with $|A|\leq n/2\log^5 n$, $|N^+(A,W)|\geq |A|\log^5 n$, and $|N^-(A,W)|\geq |A|\log^5 n$.
\item Any two disjoint subsets $A,B\subset V(G)$ with $|A|,|B|\geq n/2\log^5 n$ must have a directed edge from $A$ into $B$.
\end{enumerate}
Then we can cover $G$ with $n/k$ paths $P_i$, length $k-1$, so that, for each $i$, $P_i$ is a directed path from $x_i$ to $y_i$.
\end{theorem}
\subsection{Probabilistic results}
We will use the following results to get expansion properties in a random graph.
\begin{prop}\label{generalprops} Almost surely, if $np>20$, any two disjoint subsets $A,B\subset V(G)$ of $G=\mathcal{G}(n,p)$ with $|A|=|B|=\lceil 5\log(np)/ p\rceil$ have some edge between them.
\end{prop}
\begin{proof}
Let $m=\lceil 5\log(np)/ p\rceil$. If $q$ is the probability that there exist two disjoint subsets of size $m$ which have no edge between them, then
\[
q\leq \binom{n}{m}^2(1-p)^{m^2}\leq \left(\frac{en}{m}\right)^{2m}e^{-pm^2}
\leq \left(\frac{2enp}{5\log(np)}\right)^{2m}e^{-5m\log(np) }.
\]
Therefore,
\[
q\leq (np)^{2m}e^{-5m\log(np)}= e^{-3m\log(np)}.
\]
Now, $m\log(np)\geq 5\log^2(np)/p\to \infty$ as $n\to\infty$. Therefore, $q\to 0$ as $n\to\infty$.
\end{proof}
\begin{prop}[Alon, Krivelevich and Sudakov~\cite{AKS07}, Proposition 3.2]\label{AKS1}
Let $G=\mathcal{G}(n,p)$ be a random graph with $np>20$. Then almost surely
the number of edges between any two disjoint subsets of vertices $A$, $|A|=a$, and $B$, $|B|=b$, with $abp\geq 32n$ is at least $abp/2$ and at most $3abp/2$.
\end{prop}
Our proofs of Theorems \ref{combshort} and \ref{comblong} use ideas from the sharp embedding result of Hefetz, Krivelevich, and Szab\'o~\cite{HKS12}, and we will use the following lemmas from their work.
\begin{lemma}[\cite{HKS12}, Lemma 2.1] \label{HKS}Let $0<\varepsilon<1$ and $0\leq \beta\leq \varepsilon/7$ be real numbers and let $p=p(n)=(1+\varepsilon)\log n/ n$. Let $U\subset[n]$ have size $|U|\leq \beta n$. Then, almost surely, the random graph $G=\mathcal{G}(n,p)$ with $V(G)=[n]$ satisfies the following properties:
\begin{enumerate}
\item $\Delta(G)\leq 10\log n$.
\item $d_G(u,[n]\setminus U)\geq \eta \log n$ for every $u\in [n]$, where $0<\eta =\eta(\varepsilon)<1/2$ is a real number.
\end{enumerate}
\end{lemma}
\begin{lemma}[\cite{HKS12}, Lemma 2.4]\label{CLL} Let $G$ be a graph on $n$ vertices with maximum degree $\Delta$. Let $Y\subset V(G)$ be a set of $m=a+b$ vertices where $a$ and $b$ are positive integers. Assume that $d_G(v,Y)\geq \delta$ holds for for every $v\in V$. If
\[
\Delta^2\cdot\left\lceil\frac{m}{\min\{a,b\}}\right\rceil\cdot2\cdot e^{1-\frac{\min\{a,b\}^2}{5m^2}\cdot \delta}<1,
\]
then there exists a partition $Y=A\cup B$ of $Y$ such that
\begin{enumerate}
\item $|A|=a$ and $|B|=b$.
\item $d_G(v,A)\geq \frac{a}{3m}d_G(v,Y)$ for every $v\in V$.
\item$d_G(v,A)\geq \frac{b}{3m}d_G(v,Y)$ for every $v\in V$.
\end{enumerate}
\end{lemma}
We will also require the following lemma, which can be proved using a standard expectation argument, similar, for example, to calculations in the proof of Lemma \ref{HKS}.
\begin{lemma} \label{mindegcond} Suppose $A\subset [n]$ and $p=p(n)$ satisfy $p|A|\geq 10\log n$. Let $d=p|A|/2$. Then almost surely the random graph $G=\mathcal{G}(n,p)$ with $V(G)=[n]$ satisfies the following. For every subset $U\subset V(G)$ with $|U|\leq |A|/2d$, $|N(U,A)|\geq d|U|$.
\end{lemma}
\subsection{An important property}
The following graph property allow us to translate minimum degree conditions into expansion conditions.
\begin{defn} A graph $G$ has the $(d,D,r)$-property if it contains no sets $A,B\subset V(G)$ with $|A|\leq r$, $|B|\leq d|A|$ and $d_G(A,B)\geq D|A|$.
\end{defn}
We will typically use the $(d,D,r)$-property in the following manner. Suppose $G$ has this property and $B\subset V(G)$. Suppose further we have a set $A\subset V(G)$, with $|A|\leq r$, each of which has at least $D$ neighbours in $B$. For any subset $U\subset A$, $d(U,U\cup N(U,B))\geq D|U|$, and so, by the $(d,D,r)$-property, we must have $|N(U,B)|\geq (d-1)|U|$. That is, the subsets of $A$ expand into $B$. If, in addition, $A$ and $B$ are disjoint, then the above argument shows that $|N(U,B)|\geq |U|$ for all $U\subset A$. As Hall's generalised matching condition is satisfied, a $d$-matching from $A$ into $B$ must exist.
The following lemma, concerning when this property holds, is proved using straightforward probability, but its careful application is crucial in reaching the sharp threshold. It follows a section of the proof by Alon, Krivelevich and Sudakov of Lemma 3.1 in \cite{AKS07}.
\begin{lemma} \label{mindegexp} Suppose $p=p(n)$ and $d=d(n)$ satisfies $\log^{10}n/ n\geq p\geq 4/n$ and $d\geq 4$. Let $\alpha,\beta>0$ satisfy
\[
\alpha\log\left(\frac{\alpha}{2e\beta}\right)\geq 100.
\]
Then $G=\mathcal{G}(n,p)$ almost surely has the $(d,\alpha d\log\log n,\max\{n/d,\beta\log\log n/ p\})$-property.
\end{lemma}
\begin{proof}
If $G$ does not have the $(d,\alpha d\log\log n,\max\{n/d,\beta\log\log n/ p\})$-property, then there must exist two sets $A,B\subset V(G)$, where $|A|\leq \beta\log\log n/ p$, $|B|=d|A|$ and $d_{G}(A,B)\geq D|A|$, for $D=\alpha d\log\log n$ (adding vertices to $B$ if necessary to get equality). Let $p_r$ be the probability no two such sets occur with $|A|=r\leq\max\{n/d,\beta\log\log n/ p\}$. Bearing in mind that some of the edges might be counted twice if $A$ and $B$ overlap, we have
\begin{align*}\allowdisplaybreaks
p_r &\leq \binom{n}{r}\binom{n}{dr}\binom{dr^2}{Dr/2}p^{Dr/2}
\\ &\leq\left(\frac{en}{r}\left(\frac{en}{dr}\right)^{d}\left(\frac{2edrp}{D}\right)^{D/2}\right)^r
\\ &\leq\left(\left(\frac{n}{r}\right)^{2d}\left(\frac{2edrp}{D}\right)^{D/2}\right)^r\displaybreak[4]
\\ &\leq\left(\left(\frac{2ednp}{D}\right)^{2d}\left(\frac{2edrp}{D}\right)^{D/2-2d}\right)^r
\\ &\leq\left(\log^{20d}n\left(\frac{2edrp}{D}\right)^{D/4}\right)^r.
\end{align*}
If $r<\log n$, then $2edrp/D\leq \log^{11}n/ n$ for sufficiently large $n$, and hence, as $D=\alpha d\log\log n$, $p_r<n^{-2}$. If $r\geq \log n$, then, as $r\leq \beta\log\log n/ p$, we have
\[
p_r\leq \log^{20dr}n\left(\frac{2e\beta}{\alpha}\right)^{(\alpha dr\log\log n)/ 4}\leq \log^{(20d-100d/4)r}n\leq n^{-2}.
\]
Therefore, by looking at the sum of the probabilities $p_r$, we see the probability such a pair $A$, $B$ exists is at most $n^{-1}$.
\end{proof}
\subsection{Dividing trees}
For Theorem \ref{combshort}, we wish to find, in a tree with lots of teeth, a much smaller subtree which still has plenty of teeth. The following is a slight generalisation of a lemma by Hefetz, Krivelevich and Szab\'o~\cite{HKS12}, and is proved below.
\begin{lemma} \label{splittree}
For any $\varepsilon>0$ there exists $\beta=\beta(\varepsilon)>0$ and $n_0=n_0(\varepsilon)\in \mathbb{N}$ such that the following holds. For every tree $T$, with $|T|=n\geq n_0$, and subset $L\subset V(T)$, we can find subtrees $S, T_1, T_2\subset T$ covering $T$ so that $|S|\leq \varepsilon n$, $S$ contains at least $\beta|L|$ vertices in $L$, and $T_1$, $T_2$ are disjoint and each intersects $S$ in exactly one vertex.
\end{lemma}
\begin{corollary}\label{splitspines}
For any $\varepsilon>0$ there exists $\beta=\beta(\varepsilon)>0$ and $n_0=n_0(\varepsilon)\in \mathbb{N}$ such that the following holds for any $l,k\in \mathbb{N}$. For every tree $T$ with $n\geq n_0$ vertices and $l$ teeth length $k$, we can find subtrees $S, T_1, T_2\subset T$ covering $T$ so that $|V(S)|\leq \varepsilon n$, $S$ has at least $(\beta l-1)$ teeth length $k$ which are also teeth in $T$, and $T_1$, $T_2$ are disjoint and each intersects $S$ in exactly one vertex.
\end{corollary}
\begin{proof}
Taking the set $L$ to be the leaves at the end of the teeth length $k$ in $T$, we apply Lemma \ref{splittree}. If the tree produced, $S$, satisfies $|V(S)\cap L|\geq 2$ then, as it is connected, it must contain all the teeth with leaves in $V(S)\cap L$.
\end{proof}
\begin{defn}
Where $S$ is a tree, we will say two subtrees $S_1$ and $S_2$ \emph{divide $S$} if they cover $S$ and intersect on precisely one vertex.
\end{defn}
\begin{prop}\label{divide}
Given a tree $S$ we can find two trees $S_1$ and $S_2$ which divide $S$ for which $|S_1|, |S_2|\geq |S|/3$.
\end{prop}
\begin{proof}
Take two subtrees $S_1$ and $S_2$ which divide $S$ so that $||S_1|-|S_2||$ is minimized. Let $V(S_1)\cap V(S_2)=\{v\}$.
Suppose, without loss of generality, that $|S_1|>|S_2|$. If $|S_2|\geq |S_1|-1$ then we are done, so suppose otherwise. If $v$ has only one neighbour in $S_1$, $x$ say, then the two trees on the vertex sets $V(S_1)\setminus\{v\}$ and $V(S_2)\cup\{x\}$ intersect only on $x$ and cover $S$, contradicting the choice of $S_1$ and $S_2$.
Therefore, there must be at least two neighbours of $v$ in $S_1$, and hence we can find two subtrees $S_3,S_4\subset S_1$ which cover $S_1$, each have at least two vertices and which intersect only on $v$. Without loss of generality suppose that $|S_3|\geq |S_4|$, so that $|S_4|\leq 1+(|S_1|-1)/2$. The trees on vertex sets $(V(S_1)\setminus V(S_4))\cup\{v\}$ and $V(S_2)\cup V(S_4)$ divide $S$, so, to avoid contradicting the choice of $S_1$ and $S_2$, we must have
\[
|S_1|-|S_2|\leq |S_4|-1\leq (|S_1|-1)/2.
\]
Therefore, $2|S_2|\geq |S_1|$ and so, as $|S_2|+|S_1|=|S|+1$, $|S_2|\geq |S|/3$.
\end{proof}
\begin{proof}[Proof of Lemma \ref{splittree}] We will prove the lemma for $\varepsilon=(3/4)^{-k}$ with parameters $\beta=6^{-k}$ and $n_0=12(4/3)^k$ by induction on $k\in \mathbb{N}\cup\{0\}$. This will prove the lemma for all $\varepsilon>0$ by taking some integer $k=k(\varepsilon)$ such that $(3/4)^{-k}<\varepsilon$. The statement holds easily for $k=0$.
Suppose then the statement holds for $k$. Given a tree $T$ with at least $12(4/3)^{k+1}$ vertices and a set $L\subset V(T)$ with $l=|L|$, find the trees $S$, $T_1$ and $T_2$ as described by the lemma for $k$ and say that $V(T_1)\cap V(S)=\{t_1\}$ and $V(T_2)\cap V(S)=\{t_2\}$. Note that if $|S|\leq (3/4)^{k+1}|T|$, then we are done, so suppose $|S|\geq (3/4)^{k+1}|T|\geq 12$. Divide $S$ into the subtrees $S_1$ and $S_2$ using Lemma \ref{divide}, so that $|S_1|,|S_2|\geq |S|/3$. Then $|S_1|,|S_2|\leq 2|S|/3+1\leq (3/4)^{k+1}|T|$. As $|S\cap L|\geq 6^{-k}|L|$, without loss of generality, we have $|S_1\cap L|\geq 6^{-k}|L|/2$. Let $V(S_1)\cap V(S_2)=\{s\}$. Note that possibly $s\in\{t_1,t_2\}$.
Suppose $t_1$, $t_2$, $s$ are distinct vertices which all lie in $S_1$. Take the common intersection vertex of the unique $t_1,t_2$-path, the $t_2,s$-path and the $t_1,s$-path in $S_1$, and call it $u$. Deleting $u$ disconnects the vertices $t_1$, $t_2$ and $s$ from each other, so we may find trees $S'_1$, $S'_2$ and $S'_3$ which intersect only on $u$, contain $t_1$, $t_2$ and $s$ respectively and cover $S$. One of these trees, $S'_j$ say, must satisfy $|S'_j\cap L|\geq 6^{-(k+1)}|L|$. Then, $S_j'$ shares at most two other vertices with the other trees $S'_i$ and the trees $S_2$, $T_1$, $T_2$, so we may proceed as below.
If the tree $S_1$ shares at most two other vertices with the trees $S_2$, $T_1$ or $T_2$ then merging any of the trees $S_2$, $T_1$ and $T_2$ which share a vertex gives at most two trees, each of which intersects with $S_1$ on precisely one vertex. Taking an additional tree consisting of a single vertex of $S_1$ if necessary, to make up a second tree, we have the induction hypothesis for $k+1$.
\end{proof}
\subsection{Embedding $\mathrm{Comb}_{n,\sqrt{n}}$ in $\mathcal{G}(n,\log^2n/ n)$}
\begin{lemma}\label{log2sketch} If $\sqrt{n}\in \mathbb{N}$, then almost surely there is a copy of $\mathrm{Comb}_{n,\sqrt{n}}$ in $G=\mathcal{G}(n,\log^2n/ n)$
\end{lemma}
\begin{proof}[Sketch proof of Lemma \ref{log2sketch}]
We reveal the edges of $G$ in three rounds, where at each stage any edge is present independently with probability $p=\log^2 n/3n$. The edges of the final graph appear then with probability at most $1-(1-p)^2\leq 3p$. Therefore, if a copy of $\mathrm{Comb}_{n,\sqrt{n}}$ exists almost surely in such a random graph, then almost surely one must exist in the random graph $\mathcal{G}(n,\log^2 n /n)$.
Reveal the first set of edges to get $G_1$. Almost surely, any subgraph $H\subset V(G)$ with $|H|\geq n/3$ will satisfy the conditions of Lemma \ref{rotatingpaths} with $\lambda=1$ (using, for example, Proposition \ref{AKS1}). Therefore, any subset of $n/3$ vertices must contain an $(l,\gamma)$-connector for $\gamma=\log\log n/16\log n$ and $l=\lceil(\sqrt{n}-1)/2\rceil$. We may then find greedily in the graph a set $\mathcal{A}$ of $\sqrt{n}$ disjoint $(l,\gamma)$-connectors. Find a path length $\sqrt{n}-1$ and label it $Q=q_1\ldots q_{\sqrt{n}}$ (for example by taking a path through a $(\sqrt{n},\gamma)$-connector). This path will form the `spine' of our comb.
Let $W$ be the set of vertices in the graph not in any of the connectors or the path $Q$, so that $|W|=\sqrt{n}\lfloor(\sqrt{n}-1)/2\rfloor\geq n/3$. By revealing more edges with probability $p$ to get $G_2$ we can almost surely find a Hamilton cycle in $G_2[W]$. Take this cycle and break it into $\sqrt{n}$ paths length $\lfloor(\sqrt{n}-1)/2\rfloor-1$. Label these paths by $R_i$, $i\in[\sqrt{n}]$, and in each path pick an end vertex and label it $r_i$.
We have now covered our graph by a path of length $\sqrt{n}-1$, $Q$, the $\sqrt{n}$ $(\lceil(\sqrt{n}-1)/2\rceil,\gamma)$-connectors in $\mathcal{A}$, and $\sqrt{n}$ paths of length $\lfloor(\sqrt{n}-1)/2\rfloor-1$, $R_i$, as illustrated in Figure \ref{absorbpic}. Suppose there is a matching in the bipartite grouped graph on vertex sets $\{q_i:i\in[\sqrt{n}]\}$ and $\mathcal{A}^+$, and a matching in the bipartite grouped graph on vertex sets $\mathcal{A}^-$ and $\{r_i:i\in[\sqrt{n}]\}$. Then, for each $i$, we could take the connector $P\in \mathcal{A}$ for which $P^+$ is matched to $q_i$, and take the vertex $r_j$ matched to $P^-$, and find a neighbour $q$ of $q_i$ in $P^+$ and a neighbour $r$ of $r_j$ in $P^-$. Taking a $q,r$-Hamilton path in $G_1[P]$, we could then attach a tooth length $\sqrt{n}-1$ to $q_i$ using this path and $R_j$ along with the edges $q_iq$ and $rr_i$. Doing this for each $i\in[\sqrt{n}]$ gives a copy of $\mathrm{Comb}_{n,\sqrt{n}}$.
These matchings almost surely exist if we reveal more edges with probability $p$. Indeed, the probability an edge in the two grouped graphs is present is at least
\[
1-(1-p)^{\gamma l}\geq p\gamma l/2\geq \frac{\log n\log\log n}{16\sqrt{n}}.
\]
Thus the two grouped graphs are random bipartite graphs with equal class sizes $m=\sqrt{n}$ and edges present independently with probability at least $2\log m /m$. This is above the threshold for a matching to almost surely exist in such a graph (see, for example, Bollob\'{a}s~\cite{bollo1}).
\end{proof}
\input{sharppicture.tex}
The limiting requirement for the probability used in this sketch is in finding the matchings, which is limited by the need to ensure that every vertex $q_i$ or $r_i$ has some neighbour in a set $P^+$ or $P^-$ respectively for some connector $P\in \mathcal{A}$. If the probability $p=C\log n/ n$, for any constant $C$, is used then we must expect vertices which have no neighbours in the sets $P^+$ or $P^-$ for any $P\in \mathcal{A}$. However, on average vertices will still have many such neighbours, at least $c\log\log n$, for some small $c$. If we can construct the paths $Q$ and $R_i$ so that all the vertices $r_i$ and $q_i$ have above the average of such neighbours then, using Lemma \ref{mindegexp}, finding the matchings will be easier. The proof of Theorem \ref{combshort} is more involved, but this is the basic idea behind reducing the probability required.
\subsection{Constructing $(l,\gamma)$-connectors}
Our construction of $(l,\gamma)$-connectors is inspired by the celebrated technique of P\'osa rotation, introduced by P\'osa to find Hamilton cycles in random graphs~\cite{posa76}. Though we do not use P\'{o}sa rotations explicitly, essentially we construct our $(l,\gamma)$-connector as a path with additional edges that guarantee the path can be rotated at both ends, independently, to give at least $\gamma l$ new end vertices.
\begin{proof}[Proof of Lemma \ref{rotatingpaths}]
Let $k=4\log n / \log\log n$, $m=\lfloor\lambda n/\log n\rfloor$, and $d=n/200m\geq \log n/200\lambda$. Assume that $l_0\geq 2k$, otherwise we may simply take a path length $l_0$ (found using, for example, Lemma \ref{connect}) as an $(l_0,\gamma)$-connector.
We will call a path $P$ starting at $p_0$ a \emph{good path} if there is a set $R\subset E(P)$ such that
\begin{itemize}
\item $|R|\geq |P|/ 2k$, and,
\item for each $e\in R$, there is a Hamilton path in $G[V(P)]$ which starts at $p_0$, passes through each edge in $R$ and whose last edge is $e$.
\end{itemize}
We say that such a set $R$ \emph{demonstrates that $P$ is a good path}. Let $S$ be any set in $G$ containing at least $n/3$ vertices.
\begin{claim}
Suppose we have a collection of disjoint good paths $P_i$, $i\in I$, in $G[S]$ with initial vertices $p_i$, such that
\begin{equation}\label{pathprop}
2m\leq \sum_i\left\lceil \frac{|P_i|}{2k}\right\rceil\quad\quad\text{and}\quad\quad\sum_i|P_i|\leq \frac{n}{6}.
\end{equation}
Suppose also we have a set of integers $k_i$, $i\in I$, with $k-1\leq k_i\leq 2k-1$. Then
we can find a new path $P'_j$, for some $j\in I$, with initial vertex $p_j$, which is a good path in $S$, is disjoint from all the paths $P_i$, $i\neq j$, and contains $k_j$ more vertices than $P_j$ does.
\end{claim}
In other words, given such a collection of paths satisfying (\ref{pathprop}) we can lengthen one of the paths using vertices from $S$ so that it is still a good path with the same starting vertex, keeping the paths disjoint.
\begin{proof}[Proof of the claim]
Suppose we have such a collection of disjoint good paths $P_i$, $i\in I$. For each path $P_i$, find a set $R_i$ which demonstrates that it is a good path. By (\ref{pathprop}), $\sum_i|R_i|\geq \sum_i\lceil |P_i|/2k\rceil \geq 2m$. Let $S'=S\setminus(\cup_iV(P_i))$, $R=\cup_iR_i$ and $r=|R|$. Label the vertices which appear in the edges in $R$, so that $R=\{x_iy_i:1\leq i\leq r\}$ and so that if $x_iy_i\in R_j$ then $x_i$ appears earlier in the path $P_j$ starting from $p_j$ than $y_i$ does. This labelling ensures that the vertices $x_i$ are distinct and that the vertices $y_i$ are also distinct. For each $i$, find $j$ for which $x_iy_i\in R_j$ and let $k'_i=k_j$.
By Lemma \ref{connect}, as $|R|\geq 2m$ and $|S'|\geq n/6$, for some $i$ there is an $x_i$,$y_i$-path $Q$ of length $k'_i+1$ whose interior vertices lie in $S'$. Say that the edge $x_iy_i$ is in the path $P_j$, so that $Q$ has length $k'_i+1=k_j+1$.
Let $P_j'$ be the path formed by replacing the edge $x_iy_i$ in $P_j$ by the path $Q$. This lengthens $P_j$ by the correct number of vertices, and, to finish the proof of the claim, we need only show that $P_j'$ is a good path. We will do this by finding a demonstrating set.
As $P_j$ is a good path, there is a path $P'$ through $V(P_j)$ starting at $p_j$, which passes through each edge in $R_j$ and whose final edge is $x_iy_i$. Switch the labelling of $x_i$ and $y_i$, if necessary, so that this path ends in $y_i$. Label the vertices of $Q$ so that $Q$ is the path $x_iq_1q_2\ldots q_{k_j}y_i$.
Let $R'=(R_j\setminus \{x_iy_i\})\cup\{q_1q_2,q_{k_j}y_i\}$. Then $|R'|=|R_j|+1\geq (|V(P_j)|+k_j)/2k$. Let $e\in R_j\setminus\{x_iy_i\}$. By the definition of $R_j$, there is a Hamilton path in $G[V(P_j)]$ which passes through all the edges of $R_j$ and ends in $e$. This path must contain the edge $x_iy_i$, so, by replacing that edge with the path $Q$, we get a Hamilton path in $G[V(P'_j)]$ which passes through all the edges of $R'$ and ends with $e$. Recalling the path $P'$, replace the edge $x_iy_i$ by $Q$ to get a Hamilton path in $G[V(P'_j)]$ which passes through all the edges in $R'$ and ends in $q_{k_j}y_i$. If we add instead the path $y_iq_{k_j}\ldots q_2q_1$ to the end of $P'$ then we get a Hamilton path in $G[V(P'_j)]$ which passes through all the edges in $R'$ and ends with $q_1q_2$. Thus, $R'$ demonstrates that $P'_j$ is a good path, and gives the claim.
\end{proof}
Let $l_1=\lfloor (l_0-1)/2\rfloor$. Given any collection of $2m$ edges $\{x_iy_i:i\in I\}$ and a set $S$ of at least $n/3$ vertices disjoint from these edges, we claim we can find a good path of length $l_1$, which has $x_j$ as its initial vertex, for some $j$, and which lies in $S\cup\{x_j,y_j\}$.
To show this, consider each edge $x_iy_i$ as a path $P_i$ of length 1, where the set $\{x_iy_i\}$ demonstrates that this is a good path. Thus, the edges $x_iy_i$ satisfy the conditions of the claim above.
Repeatedly apply the claim with $k_i=k-1$ if $l_1\geq |P_i|+2k-1$, and $k_i=l_1-|P_i|$ otherwise. The paths lengthen until either one of them has length $l_1$, or until the upper bound in (\ref{pathprop}) is not satisfied. In the latter case, we discard a shortest path repeatedly until the upperbound in (\ref{pathprop}) holds again. By repeatedly applying the claim, and then removing paths, eventually we must find a path length $l_1$, as required.
This claim also holds identically with $l_2=\lceil l_0/2\rceil$ in place of $l_1$.
We can now build our $(l_0,\gamma)$-connector. First, we find in the graph $G$ $4m$ disjoint paths with length 2 and label their vertices $y_ix_iy'_i$, $i\in[4m]$. If $U\subset V(G)$ satisfies $|U|\geq n/3$, then pick two disjoint subsets $U_1,U_2\subset U$ with $|U_1|=|U_2|=m$. As $|N(U_1,U)\cap N(U_2,U)|\geq |U|-n/32 -2m$, using the expansion property for $G$, we can certainly pick vertices $v\in U$, $u_1\in U_1$ and $u_2\in U_2$ so that $vu_1$, $vu_2\in E(G)$. Therefore in any subset of $G$ of size at least $n/3$ we can find a path with length 2, so we may greedily select the paths described.
Divide the vertices not in these short paths into two sets $S_1$ and $S_2$ of size at least $n/3$. Given any subset $M\subset[4m]$ of size $2m$ we can find an index $i\in M$ and a good path $P$ in $S_1\cup\{x_i,y_i\}$ with length $\lfloor (l_0-1)/2\rfloor$ which starts with $x_i$. Therefore this must be possible for at least $2m+1$ values of $i\in[4m]$. Similarly, for at least $2m+1$ values of $i$ there must be a good path from $x_i$ in $S_2\cup\{x_i,y'_i\}$ with length $\lceil (l_0-1)/2 \rceil$. There must be then some index $j\in[4m]$ and paths $P_1$ and $P_2$ which are good, start on $x_j$, are disjoint except for the vertex $x_j$, and have length $\lceil (l_0-1)/2\rceil$ and $\lfloor (l_0-1)/2\rfloor$ respectively. Let $H$ be the subgraph $G[V(P_1)\cup V(P_2)]$.
Let $R_1$ and $R_2$ be sets demonstrating that $P_1$ and $P_2$ respectively are good paths. For each $e\in R_1$ there is vertex $v$ in $e$ for which there is a Hamilton path in $G[V(P_1)]$ from $x_j$ to $v$ which goes through every edge in $R_1$. Pick such a vertex and call it $v_e$. Let $H^+=\{v_e:e\in R_1\}$. Note that if $v_e=v_{e'}$ then there is a path going through all the edges in $R_1$ which ends in $v_e$, so $e=e'$. Therefore $|H^+|\geq (l_0-1)/4k\geq \gamma l_0$. Define similarly vertices $v_e$ for each edge $e\in R_2$ and let $H^-=\{v_e:e\in R_2\}$, so that $|H^-|\geq \gamma l_0$.
For any pair of vertices $x\in H^+$ and $y\in H^-$ we may combine an $x,x_j$-Hamilton path in $G[V(P_1)]$ with a $x_j,y$-Hamilton path in $G[V(P_2)]$ and get a path from $x$ to $y$ covering exactly the vertices in $H$. Thus, $H$ is an $(l_0,\gamma)$-connector.
\end{proof}
\section{Introduction}\label{intro}
\input{1introsharp}
\section{Preliminaries}\label{tools}
\input{2tools}
\section{Almost-spanning trees}\label{almostsec}
\input{3aalmost}
\section{$(l,\gamma)$-connectors}\label{rotate}
\input{3rotate}
\section{Proof of Theorem \ref{combshort}}\label{secshort}
\input{4proofthm1}
\section{Proof of Theorem \ref{comblong}}\label{seclong}
\input{5proofthm2}
\begin{acknowledgements} The author would like to thank Andrew Thomason for his help and suggestions.
\end{acknowledgements}
\bibliographystyle{plain}
|
\section{Introduction}
The infrared limit of the QED was modeled by Bloch and Nordsieck in
1937, and their treatment of the IR singularities has become a
textbook material since. In the framework of the Bloch-Nordsieck (B-N)
model one is able to resum all of the radiative contributions to the
fermionic Green's function generated by ultra-soft photons. A detailed
discussion of this calculation can be found in the original paper of
Bloch and Nordsieck \cite{BN}, in textbooks
\cite{BogoljubovShirkov},\cite{Fried} and in \cite{Jakovac:2011aa}.
Besides an exact solution one can also give solutions in different
approximations. In particular the 2-particle-irreducible (2PI)
approximation \cite{2PIhist} is a well known tool to study the quasiparticle
properties of the system. One of the biggest challenge in front of the
2PI techniques is the representations of symmetries, in particular
gauge symmetries \cite{RS1}. The study of B-N model provides an
excellent tool to test the reliability of fixed gauge calculations.
In our paper \cite{Jakovac:2011aa} we used the 2PI functional method
to study the B-N model at zero temperature. The spectrum could be
obtained by applying numerical calculations. We found on the one hand
the disappointing fact that the fit comparison to the exact propagator
was not very promising (see \cite{Jakovac:2011aa}), on the other hand,
unlike in the old-fashioned perturbation theory, the spectrum remained
regular even in the highly infrared regime (no infrared singularity
observed at the mass shell).
At finite temperature there are several studies in the literature
\cite{IancuBlaizot1, IancuBlaizot2,Fried:2008tb} to derive the
behavior of the fermion propagator. In our paper \cite{Jakovac:2011aa}
we developed a method to reproduce the exact result using the
Ward-Takahashi identities at zero temperature (cf. also
\cite{Alekseev:1982dk}).
This could have been generalized to finite temperature in
\cite{Jakovac:2013aa}. With the help of this method we managed to
obtain a fully analytical form of the excitation spectrum. Having
these analytic results gives us a perfect opportunity to investigate
the validity of the 2PI quasiparticle description of an interacting
quantum field theory at finite temperature.
The purpose of the present paper is to show how the 2PI works at
finite temperature. We will derive the spectral function numerically
and compare it to the exactly calculated case. The upshot is: there
exists a mapping between the coupling constants of the 2PI and the
exact results in such a way that the two spectral function overlap
almost entirely. This is a highly nontrivial result, since the exact
spectral function is an asymmetric function of the frequency, rather
different from a simple Lorentzian. The most important message to the
2PI community is that our result validates the 2PI approximation
method at non-zero temperature, and only a finite reparametrization of
the theory is needed.
From the perturbative point of view the 2PI technique resums the two
particle irreducible diagrams, but the coupling constant and also the
higher point functions are remained unchanged. So for a certain 2PI
diagram there exists another infinite set of diagrams providing
coupling constant modification. In the sense of the renormalization
group we may try to take into account the sum of these diagrams
effectively as a temperature dependent (running) coupling
constant. Since we now know the value of an observable exactly (the
electron spectral function for any frequency and temperature in a
given gauge), the best method to extract the temperature dependent
coupling is to compare the 2PI and the exact results. This is done in
the present paper.
The structure of the paper is as follows. First we give an
introduction to the Bloch-Nordsieck model itself and to the
conventions of the finite temperature real time formalism. In Section
III we recap the zero temperature results: the one-loop correction
obtained from perturbation theory (PT) and the implementation of the
2PI numerics. In Section IV we derive the one-loop self energy at
$T\neq 0$ and show its consistency with the zero temperature result by
taking the $T\rightarrow 0$ limit. Then we calculate the expression
for the discontinuity of the self energy for the 2PI procedure. The
numerical implementation of the calculation happens in a similar
fashion as the zero temperature one. In Section V we present our
results obtained from the numerics and the comparison to the exact
result \cite{Jakovac:2013aa}. We found a non-trivial mapping of the
coupling between the two calculation methods from which we conclude:
the 2PI, although it is an approximation, at finite temperature it
gives a perfect qualitative description of the collective excitation
of the system.
\section{The properties of the Bloch-Nordsieck model}
The Bloch-Nordsieck model was designed to describe accurately the low
energy regime of quantum electrodynamics. Considering the
contributions to the fermion self energy only from the deep infrared
photons, reducing the Dirac spinor to a one component fermion is well
justified. Indeed, at this energy scale photons do not have enough
energy even to flip the spin of the fermion, not to mention the pair
creation \cite{BN}. In this respect, one can substitute a four vector
$u_\mu$ in the place of the $\gamma_\mu$ matrices which is considered
as the four velocity of the fermion.
The Lagrangian then reads:
\begin{eqnarray}
\mathcal{L}=-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}+\Psi^{\dagger}(iu_{\mu}D^{\mu}-m)\Psi.
\end{eqnarray}
Where the usual notations for the field-strength tensor and for the covariant derivative are used: $F_{\mu\nu}=\partial_{\mu}A_\nu-\partial_{\nu}A_\mu$ and $D_{\mu}=i\partial_{\mu}-eA_{\mu}$, respectively.
Here, as it was mentioned above, $u_{\mu}$ is a four velocity, but we can also choose the form $u=(1,\bf{v})$, with ${\bf{v}}={\bf{u}}/u_0$ by rescaling the fermionic field as $\Psi\rightarrow\Psi/\sqrt{u_0}$ and the fermion mass by $m\rightarrow m u_0$.
\par It is possible to obtain exactly the full fermion propagator associated with this theory both for zero and finite temperatures as it is presented in \cite{Jakovac:2011aa} and \cite{Jakovac:2013aa}, respectively. Now we are going to discuss the notations and conventions that we are using in this paper.\\
For the calculations we used the real time formalism (details in
\cite{Jakovac:2013aa},\cite{LeBellac}). The propagators are matrices
in this convention:
\begin{equation}
i{\cal G}_{ab}(x)=\exv{T_C \Psi_a(x) \Psi_b^\dagger (0)}\qquad
\mathrm{and}\qquad iG_{\mu\nu,ab}(x)=\exv{T_C A_{\mu a}(x) A_{\nu b} (0)},
\end{equation}
where $T_C$ denotes ordering with respect to the contour variable
(contour time ordering). At finite temperature with help of the KMS relation we can determine $G_{12}$ and $G_{21}$
\begin{equation}
\label{eq:id1}
iG_{12}(k) = \pm n_\pm(k_0) \rh(k),\qquad iG_{21}(k) = (1\pm
n_\pm)(k_0) \rh(k),
\end{equation}
where
\begin{equation}
n_\pm(k_0) = \frac1{e^{\beta k_0}\mp 1}\quad \mathrm{and}\quad \rh(k)
= iG_{21}(k)-iG_{12}(k)
\end{equation}
are the distribution functions (Bose-Einstein (+) and Fermi-Dirac (-)
statistics), and the spectral function, respectively. We will also use
R/A formalism \cite{Jakovac:2013aa}, where
\begin{equation}
\label{eq:id2}
G_{rr}=\frac{G_{21}+G_{12}}2,\quad G_{ra} = G_{11} - G_{12},\quad
\rh=i G_{ra}-iG_{ar}.
\end{equation}
The $G_{ra}$ propagator is the retarded, the $G_{ar}$ is the advanced propagator, $G_{rr}$ is usually called the Keldysh propagator.
At zero temperature the (free) fermionic Feynman-propagator reads:
\begin{equation}\label{ferm}
{\cal G}_0(p) = \frac1{u_\mu p^\mu -m +i\ep}.
\end{equation}
It has a single pole which means that there are no antiparticles in
the model. Consequently, closed fermion loops are excluded, thus there
is no self energy correction to the photon propagator at zero
temperature. Physically this means that the energy is not sufficient
to excite the antiparticles. We interpret $u^\mu$ as the fermionic
four-velocity, and since it is fixed, the soft photons cannot change
it (no fermion recoil). In fact this means that the fermion is a hard
probe of the system, hence not part of the thermal medium
\cite{IancuBlaizot1, IancuBlaizot2}. This is also supported by
the spin-statistics theorem \cite{PeshkinSchroeder} which would forbid
a one-component fermion field. Consequently we will neglect the '12'
fermion propagator, too: ${\cal G}_{12}=0$.
The exact photon propagator reads in Feynman gauge
\begin{equation}\label{phot}
G_{ab,\mu\nu}(k) = -g_{\mu\nu} G_{ab}(k),\qquad G_{ra} =
\frac1{k^2}\biggr|_{k_0\to k_0+i\ep},\quad \rh(k)= 2\pi\sgn(k_0) \delta(k^2),
\end{equation}
all other propagators can be expressed using identities \eqref{eq:id1}
and \eqref{eq:id2}.
\section{Recap of T=0 2PI calculations}
The main idea is to use the exact propagators in the perturbation
theory as building blocks of a loop integral, where the exact
propagator is determined self-consistently using skeleton diagrams as
resummation patterns. The one-loop 2PI fermion self energy diagram in the
case of the Bloch-Nordsieck model generates the resummation of all the
``rainbow'' diagrams. One needs to take care of the UV divergences,
too, on which we perform a renormalization with the same form of
divergent parts of the counterterms as in the 1-loop case.
At zero temperature we have the following self-consistent system of
equations in the 2PI approximation
\begin{eqnarray}\label{self}
\G(p)&=&\frac{1}{\G_{0}^{-1}(p)-\Sigma(p)}, \\
\label{selfloop}
-i\Sigma(p)&=&(-ie)^2\pint{4}{k}i G_{\mu\nu}(k)i\G(p-k).
\end{eqnarray}
Here $\G_0$ and $\G$ stand for the free and the dressed fermion propagator. $G_{\mu\nu}$ is the photon propagator.
\subsection{The one-loop correction}
In strict perturbation theory (PT) we use the propagators from Eq.(\ref{ferm}) and (\ref{phot}) to compute the self energy. We choose a reference frame which in $u^{\mu}=(1,0,0,0)$, and we find using dimensional regularization
\begin{equation}
\Sigma_{1loop}(p_0)=\frac{\alpha}{\pi}(p_0-m)\left(-\ln{\frac{m-p_0}{\mu}} + \cD_\epsilon \right).
\end{equation}
Here $\mu$ is the renormalization scale. The divergent part $\cD_\epsilon$ has the following expression:
\begin{equation}
\cD_\epsilon=\frac{1}{2\epsilon}+\frac{1}{2}\left(\ln{4\pi} -\gamma_E \right).
\end{equation}
We renormalized the self energy using the $\overline{\text{MS}}$ scheme, by which the counterterms read as
\begin{equation}\delta Z_{1,\overline{\text{MS}}}=\delta_{m,\overline{\text{MS}}}= \frac{\alpha}{\pi}D_\epsilon,
\end{equation}
where $\delta Z_{1,\overline{\text{MS}}}$, $\delta_{m,\overline{\text{MS}}}$ are the wave function renormalization and the multiplicative mass renormalization, respectively.
Hence, the renormalized self energy is
\begin{equation}\label{SE_1LO}
\Sigma_{1loop}^{ren}=-\frac{\alpha}{\pi}(p_0-m)\ln{\frac{m-p_0}{\mu}}.
\end{equation}
For the details see \cite{Jakovac:2011aa}.
\subsection{The 2PI procedure at $T=0$}
In the 2PI approach we treat Eq.(\ref{self}) and Eq.(\ref{selfloop}) self-consistently. Then we implement the following steps numerically \cite{Jakovac:2011aa}, which will be applied at finite T, too:\newline
\vskip 5pt \textit{step 1:} We calculate the discontinuity of the self-energy in order to use it the spectral representation of the retarded Green's function.
\begin{eqnarray}
\Sigma(p)=i e^2\pint{4}{k}G_{\mu\nu}\G(p-k)&=&ie^2\int\limits_{0}^{\infty}\frac{d\omega}{2\pi}\pint{4}{k}\frac{1}{k^2+i\epsilon}\frac{\rho(\omega)}{p_0-k_0-\omega+i\epsilon}\\
\Sigma(p)&=&\int\limits_{0}^{\infty}\frac{d\omega}{2\pi}\rho(\omega)\Sigma_{1-loop}(p,\omega)
\end{eqnarray}
Now we can take the discontinuity
\begin{eqnarray}
\Disc\limits_{p_0}\Sigma(p)=\int\limits_{0}^{\infty}\frac{d\omega}{2\pi}\rho(\omega)\Disc\limits_{p_0}\Sigma_{1-loop}(p,\omega)=
\frac{\alpha}{\pi}\int\limits_{0}^{\infty}d\omega(p_0-\omega)\rho(\omega).
\end{eqnarray}
In both equations we introduced the fermion spectral function $\rho(p)$. In our algorithm it serves as an input, which is usually the free fermion spectral function $\rho(p)=2\pi\delta(p-m)$.\newline
\vskip 5pt \textit{step 2:} Here we calculate the real part of self energy from its discontinuity. For this purpose we use the Kramers-Kronig relation:
\begin{eqnarray}
\text{Re}\Sigma(p_0,{\bf{\p}})=\int\limits_{-\infty}^{\infty}\frac{d \omega}{2 \pi} \frac{\Disc_{\omega}i\Sigma(\omega,{\bf p })}{p_0-\omega+i\epsilon}.
\end{eqnarray}
\textit{step 3:} We renormalize the real part of the self energy using the "on-mass-shell" (OMS) renormalization scheme:
\begin{eqnarray}
&\text{Re}\Sigma(p_0=m)&=0,\\
&\displaystyle\left.\frac{d \text{Re}\Sigma(p_0)}{d p_0}\right|_{p_0=m}&=0.\\
\end{eqnarray}
\vskip 5pt \textit{step 4:} From all of this information we construct the new spectral function, which reads as
\begin{eqnarray}
\rho(p)=\frac{2\text{Im}\Sigma(p)}{(p_0-m-\text{Re}\Sigma(p))^2+(\text{Im}\Sigma(p))^2}.
\end{eqnarray}
\vskip 5pt \textit{step 5:} We set the new spectral function to be our new input, and iterate this procedure till it converges.\newline
\vskip 5pt \textit{step 4+:} As an additional step we had to include a rescaling of the spectral function which was necessary to stabilize the convergence. This step is not needed at non-zero temperature.\newline
For the zero temperature case we obtained the dressed propagator for the fermion. From the analysis it turned out that this result, being an approximation, is far from the exact solution although it is infrared finite, which cannot be claimed about the PT calculation (see \cite{Jakovac:2011aa}).
\section{Non-zero temperature}
We are working in the real time formalism, hence the Green's functions
in this picture are going to have a matrix structure. We choose the
$R/A$ basis for the matrix representation to calculate the retarded
self energy. First we are going to consider the one-loop correction
then we present a derivation of the 2PI resummed spectral function at
finite temperature. To evaluate its self-consistent equations we will
use a numerical approach which is similar to what we discussed above
for the $T=0$ case. The integral equation for the retarded self
energy at non-zero temperature in Feynman gauge reads as:
\begin{eqnarray}\label{selfT}
\Sigma_{ar}(p)=ie^2\pint{4}{k} \left[ G_{rr}(k)\G_{ra}(p-k)+G_{ra}(k)\G_{rr}(p-k) \right].\end{eqnarray}
Where $G$ and $\G$ stands for the propagator of the photon and the fermion, respectively. On Fig.(\ref{diag}) we can see the pictorial representation of the fermion self energy using Feynman diagrams.
\begin{figure}[h!]
\centering
\includegraphics[scale=0.18]{SE}
\caption{The diagrammatic representation of the self energy. The wavy line corresponds to the free (here also the exact) photon propagator with a loop momentum $k$ and the double solid line is for the exact fermion propagator with momentum $p-k$. Both the polarization and the Keldysh indices are shown.}
\label{diag}
\end{figure}
Now, if we take the discontinuity we will have:
\begin{eqnarray}\label{sigmad}
\Disc\limits_{p_0}\Sigma_{ar}(p)=e^2\pint{4}{k} \left[ G_{rr}(k)\rho_{f}(p-k)+\rho_{\gamma}(k)\G_{rr}(p-k)\right].
\end{eqnarray}
Here $\rho_f$ and $\rho_\gamma$ are the spectral functions to the fermion and the photon, respectively.
In general we can express the $rr$ propagators by the spectral function combining with the distribution function of the corresponding spin statistics:
\begin{eqnarray}
\G_{rr}(p)&=&\left(\frac{1}{2}-n_{f}(p_0)\right)\rho_{f}(p),\\
G_{rr}(p)&=&\left(\frac{1}{2}+n_{b}(p_0)\right)\rho_{\gamma}(p).
\end{eqnarray}
When inserting these expressions into Eq.(\ref{sigmad}), we get
\begin{eqnarray}\label{general}
\Disc\limits_{p_0}\Sigma_{ar}(p)&=&e^2\pint{4}{k} \left[ \left(\frac{1}{2}+n_{b}(k_0)\right)\rho_{\gamma}(k)\rho_{f}(p-k)+\rho_{\gamma}(k)\left(\frac{1}{2}-n_{f}(p_0-k_0)\right)\rho_{f}(p-k)\right]=\nn
&=& e^2\pint{4}{k} \left(1+n_{b}(k_0)-n_{f}(p_0-k_0)\right)\rho_{\gamma}(k)\rho_{f}(p-k).
\end{eqnarray}
In the last step of Eq.(\ref{general}) we get the most general form of the equation, as long as we do not specify the corresponding spectral functions.
\subsection{One-loop correction at $T\neq 0$}
For the one-loop case we have to plug in the spectral function of the free theory both for the fermion and gauge fields.
By performing this substitution our equation reads as
\begin{eqnarray}\label{long}
\Disc\limits_{p_0}\Sigma_{ar}(p)&=& e^2\pint{4}{k} \left(1+n_{b}(k_0)-n_{f}(p_0-k_0)\right)2\pi\sgn {k_0}\delta(k_{0}^2-{\bf k}^2)2\pi\delta\left(u_0(p_0-k_0)-{\bf u}({\bf p}-{\bf k})-m\right)=\nn
&=&\frac{e^2}{8\pi^3}\int\limits_{0}^{\infty} \! d k {\bf k}^2 \int\limits_{-1}^{1} \! dx \frac{(2\pi)^2}{2|{\bf k}|} \left[\left(1+n_{b}(|{\bf k}|)-n_{f}(p_0-|{\bf k}|)\right)
\delta\left(u_0p_0-{\bf u}{\bf p} -u_0|{\bf k}| -|{\bf u}||{\bf k}|x-m\right)+\right.\nn
&+&\left. \left(n_{b}(|{\bf k}|)+n_{f}(p_0+|{\bf k}|)\right)\delta\left(u_0p_0-{\bf u}{\bf p} +u_0|{\bf k}| -|{\bf u}||{\bf k}|x-m\right)\right].
\end{eqnarray}
Here we introduced the variable $x$ which stands for the cosine of the angle between the two spatial three-vectors ${\bf u}$ and ${\bf k}$. For the sake of simplicity in the following we are going to use the notations $pu\equiv p_0u_0-{\bf{pu}}$ for the Minkowski product, and $k\equiv |{\bf k}|,u\equiv|{\bf u}|$ for the absolute values of the three vectors ${\bf k}$ and ${\bf u}$, respectively.\\
First we perform the angular integration for $x$:
\begin{eqnarray}\label{asd}
\Disc\limits_{p_0}\Sigma_{ar}(p)=\frac{e^2}{4\pi u} \left(\Theta(pu-m)\int\limits_{\frac{pu-m}{u+u_0}}^{\frac{pu-m}{u-u_0}} \! dk \left(1+n_{b}(k)-n_{f}(p_0-k)\right)+\Theta(m-pu)\int\limits_{\frac{m-pu}{u+u_0}}^{\frac{m-pu}{u-u_0}} \! dk \left(n_{b}(k)+n_{f}(p_0+k)\right)\right)
\end{eqnarray}
Now we are going to use the fact that the fermion in this system is a
hard probe, thus it is not part of the heat-bath \cite{IancuBlaizot1,
IancuBlaizot2}. This manifests already in Eq.(\ref{general}) in a
way that we need to set the Fermi-Dirac distribution to zero,
otherwise we would face with inconsistencies when one would try to
take the $T\rightarrow 0$ limit:
\begin{equation}\label{nofer} n_f\equiv 0 \text{ (in the B-N framework).}\end{equation}
In that case Eq.(\ref{asd}) simplifies in the following way:
\begin{eqnarray}\label{asd2}
\Disc\limits_{p_0}\Sigma_{ar}(p)=\frac{e^2}{4\pi u} \int\limits_{\frac{pu-m}{u+u_0}}^{\frac{pu-m}{u-u_0}} \! dk \left(1+n_{b}(k)\right).
\end{eqnarray}
Evaluating the integral one gets a result consistent with the $T=0$ case:
\begin{equation}
\Disc\limits_{p_0}\Sigma_{ar}(p)=\frac{e^2}{2\pi}\Theta(pu-m)(pu-m)+\frac{Te^2}{4\pi u}\ln\left(\frac{1-e^{-\beta \frac{pu-m}{u-u_0}}}{1-e^{-\beta \frac{pu-m}{u+u_0}}}\right).
\end{equation}
This gives us the desired result for the $T\rightarrow 0$ limit, namely $\Disc\limits_{p_0}\Sigma_{ar}(p)=\frac{e^2}{2\pi}\Theta(pu-m)(pu-m)$.
\subsection{Non-zero temperature calculations for the 2PI scheme}
Now we are going to derive the 2PI resummed result for the finite temperature theory. Let us consider Eq.(\ref{self}) and Eq.(\ref{selfT}). Instead of calculating the one-loop correction by inserting free propagators, we are going to use the self-consistent fermion propagator so defining a self-consistent system of integral equations. We stick to the physical picture that the fermion is not part of the thermal medium, Eq.(\ref{nofer}) . Using the calculation in Eq.(\ref{long}) we arrive to an expression for the discontinuity of the self energy for general fermion propagator:
\begin{eqnarray}
\Disc\limits_{p_0}\Sigma_{ar}(p)&=&e^2\pint{4}{k}\left(1+n_b(k_0)\right)\rho_\gamma(k)\rho_f(p-k)\nn
&=&e^2\pint{4}{k}\left(1+n_b(k_0)\right)\frac{2\pi}{2k}(\delta(k_0-k)-\delta(k_0+k))\bar\rho_f(up-uk-m).
\end{eqnarray}
Here we used the free photon propagator as above and for the general spectral function of the fermion we introduced the notation $\rho_f(p)=\bar\rho_f(up-m)$. After some algebra we find
\begin{eqnarray}
\Disc\limits_{p_0}\Sigma_{ar}(p)=\frac{e^2}{8\pi^2}\int\limits_{-\infty}^{\infty}\! dk \int\limits_{-1}^{1} \! dx k n_b(k)\bar\rho_f(w+(u_0+ux)k).
\end{eqnarray}
Here we defined $w:=up-m$, and $x$ represents the angle between the spatial parts of $k^\mu$ and $u^\mu$, so $xku$ is the scalar product of two three-dimensional vectors like in the one-loop calculation. Actually, this can be written in a more elegant, and for the numerical implementation, a more useful way. We introduce the variable $z$ as the argument of the function $\bar\rho_f$:
\begin{eqnarray}
\Disc\limits_{p_0}\Sigma_{ar}(p)&=&\frac{e^2}{8\pi^2}\frac{1}{u}\int\limits_{-\infty}^{\infty}\! dk \int\limits_{w+(u_0-u)k}^{w+(u_0+u)k}\! dz \bar\rho_f(z) n_b(k)=\frac{e^2}{8\pi^2}\frac{1}{u}\int\limits_{-\infty}^{\infty}\! dz \bar\rho_f(z) \int\limits_{\frac{z-w}{u_0-u}}^{\frac{z-w}{u_0+u}} \! dk n_b(k).
\end{eqnarray}
In the case when the length of the three-velocity tends to zero,$u\rightarrow 0$ , we have
\begin{eqnarray}\label{u0}
\Disc\limits_{p_0}\Sigma_{ar}(p_0)=\frac{\alpha}{\pi}\int\limits_{-\infty}^{\infty} dz \bar\rho_{f}(z)(p_0-m-z)(1+n_b (p_0-m-z)).
\end{eqnarray}
For $u\neq 0$
\begin{eqnarray}\label{ufin}
\Disc\limits_{p0}\Sigma_{ar}(w)=\frac{\alpha}{2\pi}\int\limits_{-\infty}^{\infty} dz \bar\rho_f(z)\frac{T}{u}
\ln\frac{1-e^{-\beta\frac{z-w}{u_0-u}}}{1-e^{-\beta\frac{z-w}{u_0+u}}}.
\end{eqnarray}
We set $m=0$, this can be done without the loss of generality since
the two expressions in Eq.($\ref{u0}$) and Eq.($\ref{ufin}$) depends
on the variable $w=up-m$ only. That means the theory is not sensitive
where the mass-shell is placed, it can be anywhere on the real line.
With these formulae we can find the solution of the self-consistent
system of equations numerically, in a similar way we did it in the
case of zero temperature ({\it{step 1 - step 5}}), but this time we
insert the finite temperature ingredients (Eqs.~\ref{u0} and \ref{ufin}).
\section{2PI results}
We are implementing the same numerical method that we used for the
zero temperature case (Section III/B); the renormalization procedure
goes in the same way. First we observe that a small thermal mass is
generated. Interestingly this thermal mass is negative, it shifts the
spectral function to the left. In the exact solution in
\cite{Jakovac:2013aa} we found a zero thermal mass, thus we can
consider it as an artifact of the 2PI approximation, which can be
incorporated into the mass, or into the notation $w=up-m$. We should
also remark that the B-N theory describes only the softest photons
while in a realistic theory the mass modification dominantly is due to
the modes with higher momenta.
\subsection{The zero velocity case}
By applying the algorithm described in Section III/B we can obtain the
spectral function derived from the 2PI approximation for the theory,
using Eq.(\ref{u0}) as the self energy input. On
Fig.~\ref{fig:varaandT} we can see the spectral function for different
coupling values and for different temperatures. The spectrum exhibits
a pole, its width is growing with increasing coupling constant and with
increasing temperature.
\begin{figure}[htbp]
\centering
\includegraphics[height=6cm]{var_a}
\hspace*{1cm}
\includegraphics[height=6cm]{var_T}\\[-2em]
\hspace*{0cm}a.)\hspace*{7cm}b.)
\caption{The coupling constant dependence of the spectral function
in the 2PI approximation (a.) at fixed temperature $T=1$, and (b.)
at fixed fixed coupling value, $\alpha=0.5$. The curves widen with
growing coupling, and growing temperature.}
\label{fig:varaandT}
\end{figure}
In the Dyson-Schwinger approach the exact spectral function can be derived
in a closed form (at least in the the zero velocity case). We wish to
compare the 2PI results to our analytical expression obtained in
\cite{Jakovac:2013aa}:
\begin{eqnarray}\label{exact}
\rho(w)=\frac{N_{\alpha}\beta\sin{(\alpha)} e^{\beta w/2}}{\cosh(\beta w)-\cos(\alpha)}\frac{1}{\left|\Gamma\left(1+\frac{\alpha}{2\pi}+i \frac{\beta w }{2 \pi}\right) \right|^2}.
\end{eqnarray}
Here we used the notation $w=p_0-m$ again and $N_{\alpha}$ is a
normalization factor. Both for the 2PI approximation and the D-S
calculation we assumed a normalization prescription, which assigned by
$\int_{w} \rho=1$ sum rule.
To check the quality of the 2PI approximation, we can compare the
resulting spectral function with the exact one. The comparison can be seen on
Fig.(\ref{2PIvsEx}).
\begin{figure}[h!]
\centering
\includegraphics[width=6cm]{2PI_vs_Ex_nosc}
\caption{Comparing the 2PI resummed spectral function to the exact
one. The solid red line is obtained from the 2PI resummation while
the dashed blue is the exact spectral function. Both of them are
at T=1 and the couplings are $\alpha_{ex}=\alpha_{2PI}=0.5$.}
\label{2PIvsEx}
\end{figure}
We can see immediately that the two spectra are not very similar. The
reason is, as we discussed in the Introduction, that the 2PI
approximation do not sum up all the diagrams, in particular the
coupling constant corrections. To improve the 2PI calculation,
therefore, we can try to take into account the resummation of these
diagrams effectively in a renormalization group inspired way, as a
temperature dependent coupling constant. We should use a
nonperturbative matching procedure, and choose that value of
$\alpha_{2PI}$ which reproduces the exact result the most
accurately. For a perfect matching not only the coupling constant, but
also the higher point functions should also be resummed. But we may
hope that the most important effect comes from the relevant couplings,
in this case from $\alpha_{2PI}$.
Therefore our strategy will be to find the best, temperature dependent
value of the coupling constant $\alpha_{2PI}$ which yields the best
match between the exact and the 2PI spectral functions. As we can see
on Fig.~\ref{fixamp}, there exist such a value, where the matching is
almost perfect. We can observe that the fit is excellent not just at the
close vicinity of the peak region, but also for much larger momentum
regime, and it can give an account also for the asymmetric form of the
exact spectral function. For asymptotically large momenta we expect
that the two curves do not agree, according to \cite{Jakovac:2011aa},
this can also be observed on Fig.~\ref{fixamp}. This result is a
strong argument in favor of the usability of 2PI technique at finite
temperature also for gauge theories.
\begin{figure}[ht]
\begin{subfigure}{.4\textwidth}
\centering
\vspace*{0.5em}
\includegraphics[height=5.85cm]{plotarmy_1}
\caption{}
\label{plotarmy1}
\end{subfigure}
\begin{subfigure}{.4\textwidth}
\centering
\includegraphics[height=6cm]{Exfit_log_b}
\caption{}
\label{ExfitLog}
\end{subfigure}
\caption{The fitting of the exact spectral function on the 2PI
spectrum in linear (a.) and logarithmic (b.) plot. We can see an
exact match at the peak and a small deviation in the
asymptotics. The fit yields $\alpha_{2PI}=0.5$ for
$\alpha_{ex}=0.293$ at $T=1$.}
\label{fixamp}
\end{figure}
Hence we can say that the coupling which takes the value of
$\alpha_{2PI}=0.5$ in the 2PI resummation at $T=1$ is equivalent to an
$\alpha_{ex}=0.293$ in the D-S calculation at the same temperature. One can also conclude that
the vertex corrections (which are absent in the 2PI self-energy
calculations) have a role to modify the value of the renormalized
coupling. In the following we are going to look for a general relation
between $\alpha_{2PI}$ and
$\alpha_{ex}$.
We can repeat the strategy above for different temperatures. In this
way we can determine a relation $\alpha_{2PI}(\alpha_{ex},T)$
(technically it is simpler to obtain $\alpha_{ex}(\alpha_{2PI},T)$ and
invert this relation). This provides the finite temperature
dependence, or finite temperature ``running'' of the 2PI
coupling constant.
We expect that for small couplings the exact and the perturbative
values agree, since the perturbation theory gives $\alpha_{ex} =
\alpha_{2PI} + {\cal O}(\alpha_{2PI}^2)$. This is indeed the case. For
larger couplings, however, the linear relation changes. Interestingly,
we can observe that two different type of functions describe the
relation between the couplings depending on the temperature. The first
type of function which gives the mapping between the to couplings is
valid in the interval $T \in [0,12.03]$. This relation can be obtained
by a one-parameter fit between the 2PI and the exact couplings, namely:
\begin{eqnarray}\label{fit}
\alpha_{2PI}=A_T(e^{\frac{\alpha_{ex}}{A_T}}-1).
\end{eqnarray}
The result is shown on Fig.~\ref{a_u0}/a., the fit parameters ($A_T$)
are listed in Tables~\ref{tab:fitpar1} and \ref{tab:fitpar2}. From
this relation we immediately see that for small $\alpha_{2PI}$ the
relation of the couplings is linear
\begin{eqnarray}\label{universal}
\alpha_{2PI}\approx \alpha_{ex}+\mathcal{O}\left(\frac{\alpha_{ex}^2}{A_T}\right).
\end{eqnarray}
This tells us that the 2PI and the exact couplings are the same for
the perturbative region, meaning that we can rely on the results
obtained by 2PI calculations in this regime. Thus if we are using
couplings which are in the order of the fine structure constant of QED
($\alpha=1/137$) for instance, one does not even have to worry about
the temperature dependence of Eq.(\ref{fit}).
\begin{figure}[htbp]
\centering
\hbox{
\includegraphics[height=8cm]{u1_fitalfa_bt}
\hspace*{1cm}
\includegraphics[height=8cm]{u1_fitalfa_at}}
\vspace*{0em}
\hspace*{0cm}a.)\hspace*{9cm}b.)
\caption{The relation between the 2PI and the exact coupling at
$u=0$ for temperatures a.) $T\in[0,12.03]$ and b.)
$T\in(12.03,\infty)$, respectively. The dashed red line indicates
the limiting function at $T=12.03$, for details see the text. }
\label{a_u0}
\end{figure}
\begin{table}[htbp]
\centering
\begin{tabular}[c]{||l||l|l|l|l|l|l|l|l|l||}
\hline
$T$ & $0.11$ & $0.25$ & $0.5$ & $1$ & $2$ & $4$ &
$7.14$ & $10$ & $12.03$ \cr
\hline
$A_T$ & $0.213$ & $0.242$ & $0.27$ & $0.305$ & $0.343$ & $0.384 $
& $0.414$ & $0.426$ & $0.429$\cr
\hline
\end{tabular}
\caption{The fit parameters in the low temperature case. The error
of the parameters is $\pm 0.001$}
\label{tab:fitpar1}
\end{table}
\begin{table}[htbp]
\centering
\begin{tabular}[c]{||l||l|l|l|l|l|l|l||}
\hline
$T$ & $20$ & $50$ & $100$ & $200$ & $500$ & $1000$ & $2000$\cr
\hline
$B_T$ & $1.03 \pm 0.003$ & $1.118 \pm 0.008$ & $1.217 \pm 0.014$ &
$1.312 \pm 0.017$ & $1.381 \pm 0.016$ & $1.38 \pm 0.012$ & $1.3
\pm 0.006$ \cr
\hline
$C_T$ & $1.107\pm 0.001$ & $1.668 \pm 0.025$ & $2.654 \pm 0.054$ &
$4.241 \pm 0.083$ & $6.937 \pm 0.105$ & $8.951 \pm 0.1$ & $9.991
\pm 0.055$ \cr
\hline
\end{tabular}
\caption{The fit parameters in the high temperature case.}
\label{tab:fitpar2}
\end{table}
From Eq.(\ref{fit}) it is obvious that the relation depends on the
temperature through the fit parameter $A_T$: this is shown in Fig.\ref{paraA_u0}/a. We can fit
the temperature dependence in the following form:
\begin{eqnarray}\label{tanh}
A_T=a(\tanh{T b})^c,
\end{eqnarray}
where $a=0.438 \pm 0.002$, $b=0.123\pm 0.01$, and $c=0.17 \pm 0.002$.
\begin{figure}[htbp]
\centering
\hbox{
\includegraphics[height=4.9cm]{ParamA_u1}
\hspace*{1cm}
\includegraphics[height=5cm]{ParamBC_u1_0}}
\vspace*{0em}
\hspace*{-1cm}a.)\hspace*{8.5cm}b.)
\caption{The running of a.) $A_T$ and b.) $B_T/C_T$ with respect to
the temperature. This latter quantity is the position of the pole
(cf. \eqref{fit2}). One can see the best matching on higher
temperatures.}
\label{paraA_u0}
\end{figure}
Let us consider the zero temperature limit:$\lim\limits_{T\rightarrow
0} A_T = 0$. This tells us that in the zero temperature limit all
$\alpha_{ex}$ corresponding to any $\alpha_{2PI}$ by Eq.(\ref{fit})
vanishes. To see this it is easier to invert the relation and then
taking the limit, i.e. $\lim_{T\rightarrow
0}A_{T}\ln(\alpha_{2PI}/A_T+1)=0$. This is consistent with the fact
that at T=0 the coupling drops out from the 2PI propagator
\cite{Jakovac:2011aa}. More precisely at $T=0$ close to the peak:
\begin{equation}
\label{eq:2piexatT0}
\G_{2PI}(w) \propto \frac{1}{w}\quad\mathrm{, while}\qquad
\G_{ex} (w) \propto \left.\frac{1}{w^{1+\frac{\alpha_{ex}}{\pi}}}\right|_{\alpha_{ex}=0}=\G_{2PI}.
\end{equation}
Therefore the diverging $\alpha_{2PI}/\alpha_{ex}$ relation does not
signal a physical singularity, it just means that in order to match
the exact theory we have to take into account also higher point functions.
The relation in \eqref{tanh} is valid up to the dimensionless
temperature $T=12.03$. Above this temperature the trend of the curves
can be seen in Fig.~\ref{a_u0}/a, namely that they are more and more
shallow for increasing temperature, changes. The
$\alpha_{2PI}(\alpha_{ex})$ curve becomes steeper and steeper as it
can be seen in Fig.~\ref{a_u0}/b. We find for small couplings the
expected universal linear relation
$\alpha_{2PI}=\alpha_{ex}+\dots$. We can also observe that the
$\alpha_{2PI}(\alpha_{ex})$ curves diverge at some limiting value of
$\alpha_{ex}$. This can also be seen from the following fit which
describes the numerically determined curve quite well:
\begin{equation}\label{fit2}
\alpha_{2PI}=\frac{\alpha_{ex}}{B_T-C_T\alpha_{ex}}.
\end{equation}
The fit parameters can be seen in Table~\ref{tab:fitpar2}. This
function has a pole at $B_{T}/C_{T}$ at each temperature. This is a
temperature dependent quantity, the running of the position of the
pole can be seen on Fig.~\ref{paraA_u0}/b.
Formula \eqref{fit2} can be interpreted from the point of view of the
scale dependence of the coupling constant. For the B-N model the one
loop running is exact \cite{Jakovac:2011aa}, and provides a Landau
pole. The value of the coupling where we find the pole is
$\alpha(\mu_0) = \frac \pi {\ln \mu/\mu_0}$. If we associate $\mu\sim
T$ for high temperatures, this would suggest that the finite
temperature dependence also exhibits a Landau-type pole at
$\alpha_{ex} \sim (\ln f T)^{-1}$. In fact, a two-parameter fit is
\begin{eqnarray}
\frac{B_T}{C_T}=\frac{d}{\ln(f T),}
\end{eqnarray}
where $d=0.576\pm 0.03$ and $f=0.035\pm 0.003$ describes the finite
temperature behavior for large temperatures.
The finite temperature running of $\alpha_{2PI}$ for fixed
$\alpha_{ex}$ can be seen in Fig.~\ref{fig:alpha2PIrunning}.
\begin{figure}[htbp]
\centering
\includegraphics[height=6cm]{2PI_running}
\caption{Finite temperature running of $\alpha_{2PI}$ for fixed
$\alpha_{ex}$. One can observe the high temperature (Landau) pole
and the $T\to 0$ divergence.}
\label{fig:alpha2PIrunning}
\end{figure}
According to our earlier analysis we can identify the following
characteristic features of this running. For small temperatures the
running of the perturbative coupling is determined by the soft IR
physics, the photon cloud. At very small temperatures seemingly we
find a divergence, but this is not a physical singularity, it just
reflects the fact that at zero temperature the 2PI approximation fails
to describe the exact spectrum for any couplings,
cf. \eqref{eq:2piexatT0}. At high temperatures the perturbative
running is the dominant effect with the association $\mu\sim T$. Again
we find a pole there which comes from the Landau pole of the
perturbative running. But again, this singularity is not a physical
one, the exact spectrum is regular for $\alpha_{ex}$ larger than the
pole value. But with 2PI calculation with the original action we
cannot reproduce this result, one would need to take into account
higher point vertices, too. Between the low temperature and high
temperature regimes there is a point where $d\alpha_{2PI}(T)/dT =0$,
in our case this is at the dimensionless temperature value
$T=12.03$. This is a ``fixed point'' of the running and loosely
determines a ``critical temperature'' separating the two physically
different temperature regimes.
\subsection{The finite velocity case}
We can repeat the same analysis for the finite velocity case, too. Since the
findings are very similar to the $u=0$ case, we just shortly overview
the results.
For the finite velocity case we obtained in our previous article
\cite{Jakovac:2013aa} the following formula in real time:
\begin{eqnarray}\label{finuprod}
\rho(t)\propto z(t)\rho_{u=0}(t;\alpha_{eff}).
\end{eqnarray}
Where we defined an effective coupling which incorporates the
information about the finite velocity
\begin{eqnarray}
\alpha_{eff}=\frac{\alpha u_{0}(1-v^2)}{2v}\ln{\frac{1+v}{1-v}},
\end{eqnarray}
and here $v=\frac{u}{u_0}$. $z(t)$ is a function of time which is defined as:
\begin{eqnarray}
z(t)=\exp\left\{ \frac{u_0(1-v^2)\alpha}{2\pi v} \int\limits_{u_0(1-v)}^{u_0(1+v)} \!\frac{ds}{s^2} \ln\frac{\sinh\pi t Ts}{(\sinh\pi Tt)^s}\right\}.
\end{eqnarray}
In momentum space the product in Eq.(\ref{finuprod}) turns into a
convolution and thus one can derive the finite velocity spectral
function only by using numerics. In the 2PI case we are going to use
the same numerical calculation that we had for the $u=0$ case, the
only difference is that this time we use the formula in
Eq.(\ref{ufin}) for the discontinuity of the self energy. The spectral
functions obtained from 2PI for different $u>0$, but fixed temperature
and coupling constant, can be seen on Fig.(\ref{varumain}).
\begin{figure}[htbp]
\centering
\includegraphics[height=6cm]{var_u}
\caption{The 2PI spectral functions with different rapidities
($\eta=\tanh^{-1}(v)$, where $v=u/u_0$) at fixed temperature $T=1$
and coupling $\alpha=0.5$. The shrinking of the width can be
observed as the velocity grows, which is the same effect that we
had for the exact solution \cite{Jakovac:2013aa}.}
\label{varumain}
\end{figure}
\begin{figure}[htbp]
\centering
\hbox{
\includegraphics[height=8cm]{u2_fitalfa_bt}
\hspace*{1cm}
\includegraphics[height=8cm]{u2_fitalfa_at}}
\vspace*{0em}
\hspace*{0cm}a.)\hspace*{9cm}b.)
\caption{The relation between the 2PI and the exact coupling for
$u=\sqrt{3}$ at temperatures a.) $T\in[0,12.03]$ and b.)
$T\in(12.03,\infty)$, respectively. The dashed red line indicates
the limiting function at $T=12.03$, for details see the text.}
\label{a_ufin}
\end{figure}
To fit the $u>0$ spectral functions we are applying exactly the same
procedure that we used for the $u=0$ case. For this purpose we choose
the value $u=\sqrt{3}$ (or $v=\sqrt{3}/2$). In Fig.(\ref{a_ufin}) we
can find the relation between the 2PI and the exact couplings and in
Table~\ref{tab:fitpar1ufin},~\ref{tab:fitpar2ufin} the corresponding
fit parameters, but this time for $u=\sqrt{3}$. For the given finite
$u$ we have almost the same picture that we had for the $u=0$ case,
just the fit parameters, $A_T, B_T$ and $C_T$, are
different. Interestingly the threshold temperature stayed at $T=12.03$
but the running of the parameter as a function of the temperature is
slightly modified. Now we have for $A_T=a\tanh(b A_T)^c$, where this
time $a=0.55\pm 0.01$, $b=0.075\pm 0.01$ and $c=0.183\pm 0.004$. For
the running of the pole we have ($B_T/C_T=d/\ln(fT)$) $d=0.623 \pm
0.04$ and $f=0.032\pm 0.003$.
\begin{table}[htbp]
\centering
\begin{tabular}[c]{||l||l|l|l|l|l|l|l|l|l||}
\hline
$T$ & $0.11$ & $0.25$ & $0.5$ & $1$ & $2$ & $4$ &
$7.14$ & $10$ & $12.03$ \cr
\hline
$A_T$ & $0.235 \pm 0.002$ & $0.266 \pm 0.002$ & $0.298 \pm 0.002$ & $0.338 \pm 0.002$ & $0.386 \pm 0.002$ & $ 0.442 \pm 0.002 $ & $0.488 \pm 0.002$ & $0.507 \pm 0.001$ & $0.517 \pm 0.001$\cr
\hline
\end{tabular}
\caption{The fit parameters in the low temperature case for $u=\sqrt{3}$.}
\label{tab:fitpar1ufin}
\end{table}
\begin{table}[htbp]
\centering
\begin{tabular}[c]{||l||l|l|l|l|l|l|l||}
\hline
$T$ & $20$ & $50$ & $100$ & $200$ & $500$ & $1000$ &
$2000$\cr
\hline
$B_T$ & $1.016 \pm 0.001$ & $1.108 \pm 0.007$ & $1.2 \pm 0.014$ &
$1.29 \pm 0.017$ & $1.371 \pm 0.0175$ & $1.389 \pm 0.0145$ & $1.35 \pm 0.009$ \cr
\hline
$C_T$ & $0.908 \pm 0.002$ & $1.422 \pm 0.023$ & $=2.275 \pm 0.048$ &
$3.629 \pm 0.075$ & $6.113 \pm 0.102$ & $8.216 \pm 0.105$ & $9.8318 \pm 0.079$ \cr
\hline
\end{tabular}
\caption{The fit parameters in the high temperature case for $u=\sqrt{3}$.}
\label{tab:fitpar2ufin}
\end{table}
\section{Conclusion}
We gave a numerical implementation of the 2PI resummation for the
fermionic spectral function in the Bloch-Nordsieck model at non-zero
temperature. In our former paper \cite{Jakovac:2013aa} we showed a
derivation of the exact spectral function in an analytic way and
obtained a closed form. Hence, this analytic formula provides us a
good basis point in the benchmarking of the 2PI. The 2PI technique,
being an approximation, cannot provide us a full solution, but we can
still compare it to the exact result.
Our first main result is that the 2PI approximation works excellently
at finite temperatures, the spectrum coming from the 2PI approximation
could be fitted to the exact spectrum with very good accuracy. The two
curves could be fitted into each other not just in the vicinity of the
peak, but also for much larger momentum interval. This demonstrates
that the 2PI resummation is in fact a physically appropriate
approximation for gauge theories, too.
Nevertheless, the 2PI and the exact results could be fitted to each
other after properly choosing the 2PI coupling
$\alpha_{2PI}(\alpha_{ex},T)$ as a function of the coupling of the
exact formula ($\alpha_{ex}$) and the temperature. For a fixed
$\alpha_{ex}$ this describes a temperature dependent ``running''
coupling constant. Our second main result is to provide this function
for the B-N model.
This temperature dependence has two distinct regimes for small and
large temperatures. At small temperatures the deep IR physics dominate
the running, the corresponding $\alpha_{2PI}(T)$ decreases with the
temperature. For high temperatures the finite temperature running is
compatible with the perturbative scale dependence with the choice
$\mu\sim T$, there $\alpha_{2PI}(T)$ grows with the temperature. At
zero temperature and at some (coupling dependent) high temperatures we
find divergences in $\alpha_{2PI}(T)$, in the high temperature case it
can be associated with the Landau-pole. But none of these poles mean
physical singularity, just the breakdown of the perturbation
theory. Between the two regimes there is a temperature, where the
temperature derivative of $\alpha_{2PI}'(T)=0$. The ``critical
temperature'' of this ``fixed point'' is in dimensionless units
$T=12.03$, this signals the limiting temperature of the soft and
perturbative domains.
The success of the 2PI method extended by a nonperturbative running of
the coupling constant encourages one to try this strategy also in case
of other (gauge) theories. The basis of the temperature running could
be the matching of a nonperturbatively (eg. in MC simulations)
determined physical quantity. Then, using temperature dependent 2PI
couplings one could perform other calculations, and give predictions
for other, numerically hardly accessible physical quantities.
\begin{acknowledgments}
The authors acknowledge useful discussions with M. Horv\'ath,
A. Patk\'os and Zs. Sz\'ep. The project was supported by the
Hungarian National Fund OTKA-K104292. This research was also
supported by the European Union and the State of Hungary,
co-financed by the European Social Fund in the framework of
T\'AMOP-4.2.4.A/ 2-11/1-2012-0001 'National Excellence Program'.
\end{acknowledgments}
|
\section{Introduction}
\label{section-intro}
In the ab-initio quantum mechanical modeling of many-particle systems, Kohn-Sham density functional theory (DFT)
\cite{hohenberg64,kohn65} achieves so far the best compromise between accuracy and computational cost, and has
become the most widely used electronic structure model in molecular simulations and material science.
Within the traditional Kohn-Sham formulation, the ground state energy and electron density of an $N$-electron
system can be obtained by minimizing the energy functional
\begin{equation}\label{ks-energy}
E_{\rm KS}\left[\{\phi_i\}_{i=1}^{N}\right] =
\int_{\mathbb{R}^3} \left( \frac{1}{2}\sum_{i=1}^{N}|\nabla\phi_i({\bf r})|^2
+ v_{\rm ext}({\bf r})\rho({\bf r}) ] \right) d{\bf r} + E_{\rm H}[\rho] + E_{\rm xc}[\rho]
\end{equation}
with respect to orbitals $\{\phi_i\}_{i=1}^{N}$ under the constraint $\int_{\mathbb{R}^3}\phi_i\phi_j = \delta_{ij}$.
Here, $\rho({\bf r})=\sum_{i=1}^{N}|\phi_i({\bf r})|^2$ is the electron density,
$v_{\rm ext}$ is the electrostatic attraction potential generated by the nuclei,
$E_{\rm H}[\rho]=\frac{1}{2}\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}\frac{\rho({\bf r})\rho({\bf r'})}
{|{\bf r}-{\bf r'}|}d{\bf r}d{\bf r'}$ is the Hartree energy that describes electron-electron Coulomb repulsion
energy by a mean field approximation, and $E_{\rm xc}[\rho]$ is the so-called exchange-correlation energy
functional that includes all the many-particle interactions.
The major drawback of DFT is the fact that the exact functional for the exchange-correlation energy is not known.
A basic model is local density approximation (LDA) \cite{kohn65,perdew81},
which is still commonly used in practical calculations.
Improvements of this model give rise to the generalized gradient approximation (GGA)
\cite{langreth80,perdew86,perdew96} and hybrid functionals \cite{becke93,lee88,stephens94}.
Although these models have achieved high accuracy for many chemical and physical systems,
there remain well-known limitations.
For example, in systems with significant static correlation \cite{helgaker00},
LDA, GGA, and also hybrid functionals underestimate the magnitude of the correlation energy.
This becomes particularly problematic for the dissociation of electron pair bonds.
A famous example is the dissociating H$_2$ molecule: the widely employed LDA, GGA,
and even hybrid models fail rather badly at describing the energy curve for dissociating H$_2$.
Many efforts have been made in order to make an appropriate ansatz for the exchange-correlation
functional and tackle this problem (e.g., \cite{fuchs05,gruning03}).
In our view, a principal deficiency of these works is the attempt to describe strong correlation within the framework of mean field approximations.
Alternatively, DFT calculations can also be based on the strongly interacting limit of the Hohenberg-Kohn density functional, denoted ``strictly correlated electrons'' (SCE) DFT \cite{gori09, seidl07}.
This approach considers a reference system with complete correlation between the electrons, and is able to capture key features of strong correlation within the Kohn-Sham framework.
The pioneering work \cite{gori10,malet12,seidl07} has shown that the SCE ansatz can describe certain model systems in the extreme strongly correlated regime with higher accuracy than standard Kohn-Sham DFT.
However, the calculations are presently limited to either one-dimensional or spherically symmetric systems.
To our knowledge, there is no SCE-DFT calculation for dissociating the H$_2$ molecule in $\mathbb{R}^3$.
In the SCE-DFT model, the repulsion energy between strongly interacting electrons is related to
optimal transport theory. Optimal transport was historically studied in \cite{monge81} to model the most economical way of moving soil from one area to another, and was further generalized in \cite{kantorovich1940, kanto42} to the Kantorovich primal and dual formulation. The goal is to transfer masses from
an initial density $\rho_A$ to a target density $\rho_B$ in an optimal way such that the ``cost'' $c(x,y)$ for
transporting mass from $x$ to $y$ is minimized (see \cite{villani09} for a comprehensive treatment).
The Coulomb repulsion energy in the SCE-DFT model can be reformulated as the optimal cost of an optimal transport problem, if we identify the marginals with the electron density divided by number of electrons, i.e., $\rho/N$, and the cost function with the electron-electron Coulomb repulsion
\begin{equation}\label{cost_Coulomb}
c_{\rm ee}({\bf r}_1,\dots,{\bf r}_N) = \sum_{1\leq i<j\leq N}\frac{1}{|{\bf r}_i-{\bf r}_j|}.
\end{equation}
For instance, for a two-electron system within the SCE-DFT framework,
the electron repulsion energy for a given single-particle electron density $\rho$ is
\begin{multline}\label{functional-sce}
V_{\rm ee}^{\rm SCE}[\rho] = \min_{\Psi} \bigg\{
\int_{\mathbb{R}^6}\frac{|\Psi({\bf r}_1,{\bf r}_2)|^2}{|{\bf r}_1-{\bf r}_2|} d{\bf r}_1 d{\bf r}_2, \\
\int_{\mathbb{R}^3}|\Psi({\bf r}_1,{\bf r})|^2d{\bf r}_1 =
\int_{\mathbb{R}^3}|\Psi({\bf r},{\bf r}_2)|^2d{\bf r}_2 = \frac{\rho({\bf r})}{2} \bigg\}. \quad
\end{multline}
Strictly speaking, the set of admissible $|\Psi|^2$'s must be enlarged to probability measures in order to allow strict correlation, which corresponds to concentration of the many-body probability density on a lower dimensional subset \cite{cotar13a}. There are several mathematical investigations of the relations between SCE-DFT and optimal transport problems, see \cite{buttazzo12,cotar13a,cotar13b,friesecke13}, but important open problems remain. To our knowledge, the functional derivative of the SCE functional \eqref{functional-sce} (alias optimal cost functional) with respect to the electron density
(alias marginal measure) is not clear from a mathematical point of view.
However, this result is crucial for deriving the Kohn-Sham equations needed in practical calculations.
Numerical algorithms for optimal transport problems are rather sparse.
Explicit solutions for the co-motion functions are known for one-dimensional and spherically symmetric problems \cite{seidl07}, but cannot be generalized to two- and three-dimensional systems. An alternative route might be the Kantorovich dual formulation of the SCE functional \cite{buttazzo12, mendl13}. In a complementary work \cite{vuckovic2014}, the H$_2$ molecule is studied using an ansatz for the dual potential, and there is a recent simulation of a one-dimensional model H$_2$ molecule using the SCE framework \cite{malet14}.
In this paper, we give a mathematical derivation of the Kohn-Sham equations for optimal transport-based DFT which is rigorous up to physically expected smoothness and continuity assumptions (section~\ref{section-ks}), provide an efficient numerical algorithm for discretising and solving the resulting optimal transport problem for the case of two electrons without restriction to radial symmetry (section~\ref{section-numerical}), and then apply this algorithm to a self-consistent DFT simulation of the H$_2$ molecule in the dissociating limit (section~\ref{section-h2}). Finally, we show both numerically and by a rigorous mathematical argument that the SCE-DFT model is accurate for the H$_2$ molecule in the dissociating limit.
\section{Preliminaries}
\label{section-pre}
Consider a molecular system with $M$ nuclei of charges $\{Z_1,\ldots,Z_M\}$,
located at positions $\{{\bf R}_1,\ldots,{\bf R}_M\}$,
and $N$ electrons in the non-relativistic setting.
The electrostatic potential generated by the nuclei is
\begin{equation*}
v_{\rm ext}({\bf r}) = -\sum_{I=1}^{M}\frac{Z_I}{|{\bf r}-{\bf R}_I|},
\quad {\bf r}\in\mathbb{R}^3.
\end{equation*}
Within the DFT framework \cite{hohenberg64,lieb83}, the ground state density and energy of
the system is obtained by solving the following minimization problem
\begin{equation}\label{min-dft}
E_0 = \min_{\rho} \bigg\{ F_{\rm HK}[\rho]+\int_{\mathbb{R}^3}v_{\rm ext}\rho,
~ \rho\geq 0,~\sqrt{\rho}\in H^1(\mathbb{R}^3),~\int_{\mathbb{R}^3}\rho=N \bigg\},
\end{equation}
where $\rho$ is the electron density and $F_{\rm HK}[\rho]$ is the so-called Hohenberg-Kohn functional \cite{hohenberg64}.
$F_{\rm HK}$ is a universal functional of $\rho$ in the sense that it does not depend on the external potential $v_{\rm ext}$.
Unfortunately, no tractable expression for $F_{\rm HK}$ is known that could be used in numerical simulations.
The standard Kohn-Sham DFT \cite{kohn65} treats the system as $N$ non-interacting electrons,
and approximates $F_{\rm HK}[\rho]$ by a summation of the kinetic energy
\begin{multline}\label{energy-ks}
T_{\rm KS}[\rho] = \inf\bigg\{ \frac{1}{2}\sum_{i=1}^N \int_{\mathbb{R}^3}|\nabla\phi_i({\bf r})|^2 d{\bf r},
~ \phi_i\in H^1(\mathbb{R}^3), \\
~ \sum_{i=1}^N|\phi_i({\bf r})|^2 = \rho({\bf r}),~\int_{\mathbb{R}^3}\phi_i\phi_j = \delta_{ij} \bigg\}, \quad
\end{multline}
the Hartree energy $E_{\rm H}[\rho]$, and an exchange-correction energy $E_{\rm xc}[\rho]$, as shown in Eq.~\eqref{ks-energy}.
Since the standard non-interacting model cannot capture the features that result from strong correlation, it is not able to simulate strongly correlated electron systems, like the H$_2$ molecule in its dissociating limit. In contrast to that, the SCE-DFT model \cite{seidl1999, SPL1999, seidl07} starts from the strongly interacting limit (semi-classical limit) of $F_{\rm HK}$,
and gives rise to the following SCE functional (see \cite{friesecke13} for a mathematical justification)
\begin{equation}\label{energy-sce}
V_{\rm ee}^{\rm SCE}[\rho] = \inf \big\{ V_{\rm ee}[\rho_N],
~\rho_N({\bf r}_1,\dots,{\bf r}_N)\geq 0,~\rho_N~{\rm is~symmetric},~\rho_N\mapsto\rho \big\},
\end{equation}
where
\begin{equation}\label{Vee-rhoN}
V_{\rm ee}[\rho_N] = \int_{\mathbb{R}^{3N}}\sum_{1\leq i<j\leq N}
\frac{\rho_N({\bf r}_1,\dots,{\bf r}_N)}{|{\bf r}_i-{\bf r}_j|} d{\bf r}_1\cdots d{\bf r}_N,
\end{equation}
and $\rho_N\mapsto\rho$ means that $\rho$ is the marginal distribution of $\rho_N$, that is to say
\begin{equation*}
\rho({\bf r}) = N\int_{\mathbb{R}^{3(N-1)}}\rho_N({\bf r},{\bf r}_2,\dots,{\bf r}_N) d{\bf r}_2\cdots d{\bf r}_N.
\end{equation*}
The minimization in Eq.~\eqref{energy-sce} is over all symmetric $N$-point probability measures $\rho_N$ which have the given single-particle density $\rho$ as marginal, and yields the minimum of the electronic Coulomb repulsion energy over all such $\rho_N$. The SCE-DFT model takes $V_{\rm ee}^{\rm SCE}[\rho]$ as the only interaction term, replacing $E_{\rm H}[\rho]+E_{\rm xc}[\rho]$ in standard Kohn-Sham DFT.
The minimization task \eqref{energy-sce} is in fact an optimal transport problem with Coulomb cost \cite{cotar13a,buttazzo12,friesecke13}, which has two alternative formulations:
the Monge formulation and the Kantorovich dual formulation.
For the Monge formulation, one uses the ansatz
\begin{equation}\label{eq-rhoN}
\rho_N({\bf r}_1,\dots,{\bf r}_N) = \frac{\rho({\bf r}_1)}{N}
\delta({\bf r}_2-T_2({\bf r}_1))\cdots\delta({\bf r}_N-T_N({\bf r}))
\end{equation}
with $T_i:\mathbb{R}^3\rightarrow\mathbb{R}^3~(i=2,\dots,N)$ the so-called co-motion functions
(also called optimal transport maps), where we use the convention $T_1({\bf r})={\bf r}$.
The above ansatz already appears on physical grounds, without reference to optimal transport theory, in \cite{seidl07}.
Since the $N$-particle distribution $\rho_N$ in \eqref{eq-rhoN} is zero everywhere except on the set
\begin{equation}\label{set-comotion}
M = \{({\bf r},T_2({\bf r}),\dots,T_N({\bf r})),~{\bf r}\in\mathbb{R}^3\},
\end{equation}
it describes a state where the location of one electron fixes all the other $N-1$ electrons through the co-motion functions $T_i$, $i=2,\dots,N$.
The co-motion functions are implicit functionals of the density, determined by the minimization problem \eqref{energy-sce} and a set of differential equations that ensure the invariance of the density under the coordinate transformation ${\bf r}\mapsto T_i({\bf r})$ \cite{seidl07}, i.e.,
\begin{equation}\label{eq-constraint-comotion}
\rho(T_i({\bf r}))dT_i({\bf r}) = \rho({\bf r})d{\bf r}.
\end{equation}
In terms of these functions, the optimal value of \eqref{energy-sce} reads
\begin{equation}\label{problem-comotion}
V_{\rm ee}^{\rm SCE}[\rho] = \frac{1}{N}\int_{\mathbb{R}^3}
\sum_{1\leq i<j\leq N}\frac{\rho({\bf r})}{|T_i({\bf r})-T_j({\bf r})|}d{\bf r}.
\end{equation}
Note that the ansatz \eqref{eq-rhoN} is not in general symmetric under exchanging particle coordinates, nevertheless,
dropping the symmetrization does not alter the minimum value of \eqref{energy-sce}.
Alternatively, one can start from the so-called Kantorovich dual formulation \cite{kanto42}.
It has been shown in \cite{buttazzo12} that the value of $V^{\rm SCE}_{\rm ee}[\rho]$
is exactly given by the maximum of this Kantorovich dual problem
\begin{equation}\label{kantorovich}
V_{\rm ee}^{\rm SCE}[\rho] = \max\bigg\{\int_{\mathbb{R}^3}u({\bf r})\rho({\bf r})d{\bf r},~
\sum_{i=1}^{N}u({\bf r}_i)\leq\sum_{1\leq i<j\leq N}\frac{1}{|{\bf r}_i-{\bf r}_j|} \bigg\}.
\end{equation}
In what follows, we denote the maximizer of \eqref{kantorovich} by $u_{\rho}$,
which is called the Kantorovich potential.
We assume that $u_{\rho}$ is unique and depends continuously on $\rho$ in the sense that
\begin{equation}\label{u_rho_continuous}
u_{\rho_j}\stackrel{*}{\rightharpoonup}u_{\rho} \quad{\rm if}\quad \rho_j\rightarrow\rho
~{\rm in}~L^1(\mathbb{R}^3).
\end{equation}
For numerical implementations, the Kantorovich dual formulation has high complexity due to the
$3N$-dimensionality of the constraints. In comparison, the Monge formulation amounts to a spectacular dimension reduction, in which the unknowns are $N-1$ maps on $\mathbb{R}^3$ instead of one function $\rho_N$ on $\mathbb{R}^{3N}$. However, for practical purposes it is currently restricted to spherically symmetric densities and one-dimensional systems, for which the constraints \eqref{eq-constraint-comotion} can be solved semi-analytically.
Our purpose is to construct an efficient numerical discretization of the Monge formulation for $N = 2$ electrons which is applicable to non-spherical systems.
\section{Kohn-Sham equations for optimal transport based DFT}
\label{section-ks}
By taking $V_{\rm ee}^{\rm SCE}$ as the only interaction term within the Kohn-Sham DFT framework,
we can obtain the ground state approximations of energy and electron density
by solving the following minimization problem
\begin{equation}\label{min-sce}
E_0 = \inf_{\Phi} \left\{ E_{\rm KS}^{\rm SCE}[\Phi],
~\phi_i\in H^1(\mathbb{R}^3),~\int_{\mathbb{R}^3}\phi_i\phi_j=\delta_{ij}\right\},
\end{equation}
where $\Phi=(\phi_1,\dots,\phi_N)$ denotes the Kohn-Sham orbitals and
\begin{equation}\label{eq-energy-sce}
E_{\rm KS}^{\rm SCE}[\Phi]=\frac{1}{2}\sum_{i=1}^{N} \int_{\mathbb{R}^3}|\nabla\phi_i({\bf r})|^2d{\bf r}
+ \int_{\mathbb{R}^3}v_{\rm ext}({\bf r})\rho_{\Phi}({\bf r})d{\bf r} + V_{\rm ee}^{\rm SCE}[\rho_{\Phi}]
\end{equation}
with $\rho_{\Phi}({\bf r})=\sum_{i=1}^N|\phi_i({\bf r})|^2$.
We shall derive the self-consistent Kohn-Sham equations for \eqref{min-sce} in this section.
The key point is to calculate the functional derivative of the SCE functional
$\delta V_{\rm ee}^{\rm SCE}[\rho] / \delta\rho$ with respect to the single particle density $\rho$,
which is the effective one-body potential coming from the interaction term $V_{\rm ee}^{\rm SCE}[\rho]$. In the derivation below, we make various plausible assumptions on the Kantorovich potential such as uniqueness, continuous dependence on the density, and differentiability at relevant points. We believe these assumptions to be correct except possibly in exceptional situations. A fully rigorous treatment without these assumptions would be desirable, but lies beyond the scope of this paper.
Note that the functional $V_{\rm ee}^{\rm SCE}[\rho]$ is not defined on arbitrary densities,
but only on those with $\int_{\mathbb{R}^3}\rho=N$. Therefore, the definition of the functional
derivative only specifies its integral against perturbations in the corresponding ``tangent space'', that is to say perturbations that have integral zero:
\begin{equation}\label{functional-derivative}
\int_{\mathbb{R}^3}\frac{\delta V_{\rm ee}^{\rm SCE}[\rho]}{\delta\rho}\cdot\tilde{\rho}
= \lim_{\varepsilon\rightarrow0}\frac{V_{\rm ee}^{\rm SCE}[\rho+\varepsilon\tilde{\rho}]
-V_{\rm ee}^{\rm SCE}[\rho]}{\varepsilon}
\quad{\rm for~all}~\tilde{\rho}~{\rm with}~\int_{\mathbb{R}^3}\tilde{\rho}=0.
\end{equation}
The following theorem indicates that the functional derivative is nothing but the Kantorovich potential
$u_{\rho}$ with an additive constant.
\begin{theorem}\label{theo-derivative}
Assume that the maximizer $u_{\rho}$ of \eqref{kantorovich} is unique and depends continuously on
the electron density $\rho$ in the sense of \eqref{u_rho_continuous}.
Then $v_{\rm SCE}[\rho]=\delta V_{\rm ee}^{\rm SCE}[\rho]/\delta\rho$ is a functional derivative of
$V_{\rm ee}^{\rm SCE}$ at point $\rho$ in the sense of \eqref{functional-derivative} if and only if
\begin{equation}\label{sce-kantorovich}
v_{\rm SCE}[\rho]=u_{\rho}+C \quad\text{for any constant } C.
\end{equation}
\end{theorem}
\begin{proof}
For any given single-particle density $\rho$ with $\int_{\mathbb{R}^3}\rho=N$,
and any perturbation $\tilde{\rho}$ with $\int_{\mathbb{R}^3}\tilde{\rho}=0$, we define
\begin{equation*}
D_{\varepsilon} = \frac{V_{\rm ee}^{\rm SCE}[\rho+\varepsilon\tilde{\rho}]
-V_{\rm ee}^{\rm SCE}[\rho]}{\varepsilon}.
\end{equation*}
For simplicity, we assume $\varepsilon>0$. We have from \eqref{kantorovich} that
\begin{equation}\label{proof-2-a}
D_{\varepsilon} = \frac{\int_{\mathbb{R}^3}u_{\rho+\varepsilon\tilde{\rho}}\,(\rho+\varepsilon\tilde{\rho})
- \int_{\mathbb{R}^3}u_{\rho}\,\rho}{\varepsilon}.
\end{equation}
Using the fact that $u_{\rho+\varepsilon\tilde{\rho}}$ and $u_{\rho}$ are maximizers of \eqref{kantorovich}
with electron density $\rho+\varepsilon\tilde{\rho}$ and $\rho$ respectively, we have
\begin{equation}\label{proof-2-b}
\int_{\mathbb{R}^3}u_{\rho}\,(\rho+\varepsilon\tilde{\rho}) \leq
\int_{\mathbb{R}^3}u_{\rho+\varepsilon\tilde{\rho}}\,(\rho+\varepsilon\tilde{\rho})
\quad{\rm and}\quad
-\int_{\mathbb{R}^3}u_{\rho}\,\rho \leq -\int_{\mathbb{R}^3}u_{\rho+\varepsilon\tilde{\rho}}\,\rho.
\end{equation}
Substituting \eqref{proof-2-b} into \eqref{proof-2-a} gives
\begin{equation}\label{proof-2-c}
\int_{\mathbb{R}^3}u_{\rho}\,\tilde{\rho} \leq D_{\varepsilon}
\leq \int_{\mathbb{R}^3}u_{\rho+\varepsilon\tilde{\rho}}\,\tilde{\rho}.
\end{equation}
Under the uniqueness and continuity assumption \eqref{u_rho_continuous},
the right-hand side of \eqref{proof-2-c} converges to the left-hand side as $\varepsilon\rightarrow 0$.
Hence, for any $\tilde{\rho}$ with $\int_{\mathbb{R}^3}\tilde{\rho}=0$,
$\lim_{\varepsilon\rightarrow 0}D_{\varepsilon}$ exists and equals $\int_{\mathbb{R}^3}u_{\rho}\tilde{\rho}$.
This together with definition \eqref{functional-derivative} leads to $v_{\rm SCE}[\rho]=u_{\rho}$.
Note that the map
$\tilde{\rho}\mapsto\int_{\mathbb{R}^3}\frac{\delta V_{\rm ee}^{\rm SCE}[\rho]}{\delta\rho}\tilde{\rho}$
is unique up to an additive constant since $\int_{\mathbb{R}^3}\tilde{\rho}=0$.
Therefore, the functional derivative viewed as a function can be modified by any additive constant $C$.
This completes the proof.
\end{proof}
The Kantorovich potential $u_{\rho}$ is related to the co-motion functions in the Monge formulation, as noted and justified in \cite{seidl07}. In what follows, we give a more mathematical derivation of this relation, which avoids the interpretation of the effective potential as a Lagrange multiplier and clarifies the relationship between the variational principle \eqref{SGS}, introduced in \cite{seidl07}, and the Kantorovich dual variational principle. Note that the interpretation of the effective potential as a Lagrange multiplier coming from the marginal constraint is heuristically correct, but difficult to make rigorous, the difficulties being related to the notorious ``$v$-representability-problem'', as will be discussed elsewhere.
\begin{theorem}\label{theo-potentials}
Let $\rho_N$ be the minimizer of the optimal transport problem \eqref{energy-sce} with given single-particle density $\rho$.
If $u_{\rho}$ is the Kantorovich potential, i.e., the maximizer of \eqref{kantorovich}, and $u_\rho$ is differentiable, then
\begin{equation}\label{v_rhoN}
\nabla u_{\rho}({\bf r}) = \nabla_{{\bf r}}c_{\rm ee}
({\bf r},{\bf r}_2,\dots,{\bf r}_N)~~\text{on}~~\mathrm{supp}(\rho_N).
\end{equation}
In particular, if $\rho_N$ is of the Monge form \eqref{eq-rhoN}, then
\begin{equation}\label{nabla-v}
\nabla u_{\rho}({\bf r}) = \nabla_{\bf r} c_{\rm ee}({\bf r},{\bf r}_2,\dots,{\bf r}_N)
\mid_{{\bf r}_2=T_2({\bf r}),\dots,{\bf r}_N=T_N({\bf r}_N)}.
\end{equation}
\end{theorem}
\begin{proof}
We first note that $V_{\rm ee}^{\rm SCE}[\rho]$ is convex. To see this, let $\rho$ be a convex combination
$(1-t)\rho_A+t\rho_B$ for some $t\in(0,1)$, and $\rho_N^A$, $\rho_N^B$ be the minimizers of \eqref{energy-sce}
corresponding to the single-particle densities $\rho_A$ and $\rho_B$, we have
\begin{equation}\label{sce-convex}
V_{\rm ee}^{\rm SCE}[\rho]\leq (1-t)V_{\rm ee}[\rho_N^A]+tV_{\rm ee}[\rho_N^B]
=(1-t)V_{\rm ee}^{\rm SCE}[\rho_A]+tV_{\rm ee}^{\rm SCE}[\rho_B].
\end{equation}
Since $V_{\rm ee}^{\rm SCE}[\rho]$ is convex, it equals its double Legendre transform.
Denoting the Legendre transform of a functional $F$ by $F^*$, we have
\begin{equation*}
V_{\rm ee}^{\rm SCE*}[v]=\max_{\rho}\left(\int_{\mathbb{R}^3}v\rho-V_{\rm ee}^{\rm SCE}[\rho]\right).
\end{equation*}
By combining the maximization over $\rho$ and minimization over $\rho_N\mapsto\rho$ in \eqref{energy-sce},
we have
\begin{equation}\label{proof-1-a}
\begin{split}
-V_{\rm ee}^{\rm SCE*}[v] &= \min_{\rho}\left(V_{\rm ee}^{\rm SCE}[\rho]-\int_{\mathbb{R}^3}v\rho\right) \\
&= \min_{\rho}\min_{\rho_N\mapsto\rho}
\left(\int_{\mathbb{R}^{3N}}c_{\rm ee}\rho_N - \int_{\mathbb{R}^3}v\rho\right) \\
&= \min_{\rho_N\mapsto\rho} \int_{\mathbb{R}^{3N}}\rho_N({\bf r}_1,\dots,{\bf r}_N)
\left(c_{\rm ee}({\bf r}_1,\dots,{\bf r}_N)-\sum_{i=1}^N v({\bf r}_i)\right).
\end{split}
\end{equation}
Then the double Legendre transform is
\begin{equation}\label{proof-1-b}
V_{\rm ee}^{\rm SCE**}[\rho]=\max_v\left(\int_{\mathbb{R}^3}v\rho-V_{\rm ee}^{\rm SCE*}[v]\right).
\end{equation}
Combining \eqref{proof-1-a}, \eqref{proof-1-b}, and the fact that $V_{\rm ee}^{\rm SCE}$ equals its double
Legendre transform results in the following variational principle
\begin{equation}\label{SGS}
V_{\rm ee}^{\rm SCE}[\rho] = \max_v\left( \int_{\mathbb{R}^3}v\rho
+ \min_{\rho_N}\int_{\mathbb{R}^{3N}}\big(c_{\rm ee}-\sum_{i=1}^N v({\bf r}_i)\big)\rho_N \right).
\end{equation}
Note that the constraint $\rho_N\mapsto\rho$ has been eliminated in \eqref{SGS}.
For any fixed $v$, let $\mathcal{V}({\bf r}_1,\dots,{\bf r}_N)=\sum_{i=1}^N v({\bf r}_i)$.
The inner variational principle of \eqref{SGS} reads
\begin{equation*}
\min_{\rho_N}\int_{\mathbb{R}^{3N}}(c_{\rm ee}-\mathcal{V})\rho_N.
\end{equation*}
Since $c_{\rm ee}-\mathcal{V}$ is a pure multiplicative operator, it follows (provided $v$ is differentiable) that the support of
any minimizer $\rho_N$ must be contained in the set of absolute minimizers of $c_{\rm ee}-\mathcal{V}$.
Note that on the latter set, $\nabla(c_{\rm ee}-\mathcal{V})=0$. Therefore, we have
\begin{equation}\label{proof-1-c}
\nabla_{{\bf r}_i}(c_{\rm ee}-\mathcal{V}) = \nabla_{{\bf r}_i}c_{\rm ee}({\bf r}_1,\dots,{\bf r}_N)
- \nabla_{{\bf r}_i}v({\bf r}_i) = 0 \quad\text{on }\mathrm{supp}(\rho_N)
\end{equation}
for $i=1,\dots,N$.
According to Lemma~\ref{lemma-outerSGS} in the appendix, if $v_0$ is a maximizer of \eqref{SGS},
then the corresponding minimizer $\rho_N^0$ of the inner optimization of \eqref{SGS}
is exactly the minimizer of the original problem \eqref{energy-sce}.
Therefore, \eqref{proof-1-c} implies that if $\rho_N^0$ is a minimizer of \eqref{energy-sce}, and $v_0$ is differentiable, then
\begin{equation}\label{proof-1-z}
\nabla_{{\bf r}_i}v_0({\bf r}_i) = \nabla_{{\bf r}_i}c_{\rm ee}({\bf r}_1,\dots,{\bf r}_N)
\quad{\rm on~supp}(\rho_N^0),~~i=1,\dots,N.
\end{equation}
In order to obtain \eqref{v_rhoN}, it is now only necessary to show that $u_{\rho}({\bf r})=v_0({\bf r})+\mu$ with some constant $\mu$. Note that the maximum value of \eqref{SGS} is invariant under changing $v$ by an additive constant, because the two integrals involving $v$ cancel. Therefore, the maximization over $v$ in \eqref{SGS} may be restricted to $v$'s with the additional property
\begin{equation}\label{proof-1-d}
\min_{({\bf r}_1,\dots,{\bf r}_N)\in\mathbb{R}^{3N}}\left(
c_{\rm ee}({\bf r}_1,\dots,{\bf r}_N) - \sum_{i=1}^N v({\bf r}_i) \right)= 0.
\end{equation}
For these $v$'s, the minimization over $\rho_N$ in \eqref{SGS} can be carried out explicitly (just place the the support of $\rho_N$ at the global minimizers of $c_{\rm ee}-\mathcal{V}$).
It then follows from \eqref{SGS} that
\begin{equation}\label{proof-1-e}
V_{\rm ee}^{\rm SCE}[\rho] = \max\left\{\int_{\mathbb{R}^3}v\rho,~v~{\rm satisfies}~\eqref{proof-1-d}\right\}.
\end{equation}
Since $\int_{\mathbb{R}^3}(v + C)\rho$ is increasing as a function of the additive constant $C$, condition \eqref{proof-1-d}
can be changed into the inequality
\begin{equation}\label{proof-1-f}
\sum_{i=1}^{N}v({\bf r}_i)\leq\sum_{1\leq i<j\leq N}\frac{1}{|{\bf r}_i-{\bf r}_j|}
\end{equation}
without affecting the maximal value in \eqref{proof-1-e}. This yields the Kantorovich dual form \eqref{kantorovich}.
Therefore, any maximizer of \eqref{SGS} satisfies that $u_{\rho}({\bf r})=v_0({\bf r})+\mu$ with some constant $\mu$,
which together with \eqref{proof-1-z} implies \eqref{v_rhoN}.
If the minimizer of \eqref{energy-sce} is of the Monge form \eqref{eq-rhoN},
then we have ${\rm supp}(\rho_N)\subset M$ with $M$ given by \eqref{set-comotion}.
This implies \eqref{nabla-v} and completes the proof.
\end{proof}
Since the Coulomb cost function $c_{\rm ee}$ is given by \eqref{cost_Coulomb}, Eq.~\eqref{nabla-v} is reduced to the following relation according to Theorem~\ref{theo-potentials} (see also \cite{seidl07})
\begin{equation}\label{eq-NbodyOT}
\nabla u_{\rho}({\bf r}) = -\sum_{i=2}^N\frac{{\bf r}-T_i({\bf r})}{|{\bf r}-T_i({\bf r})|^3}.
\end{equation}
In case $N=2$, there is only one co-motion function $T$, and
\begin{equation}\label{eq-2bodyOT}
\nabla u_{\rho}({\bf r}) = -\frac{{\bf r}-T({\bf r})}{|{\bf r}-T({\bf r})|^3}.
\end{equation}
Note that solving this equation for $T({\bf r})$ gives an instance of the celebrated Gangbo-McCann formula \cite{GangboMcCann1996} for the optimal map in terms of the Kantorovich potential. For the Coulomb cost this formula takes the form \cite{cotar13a}
\begin{equation}\label{eq-GangboMcCannFormula}
T({\bf r}) = {\bf r} + \frac{\nabla u_{\rho}({\bf r})}{|\nabla u_\rho({\bf r})|^{3/2}}.
\end{equation}
However, unlike \eqref{eq-GangboMcCannFormula}, formula \eqref{eq-2bodyOT} generalizes (in the form of \eqref{eq-NbodyOT}) to many-body or multi-marginal problems. Thus formula \eqref{eq-NbodyOT} should be viewed as the correct generalization of the Gangbo-McCann formula to multi-marginal problems. We note that our derivation did not make use of Coulombic features of the cost; the same arguments yield a version of Eq.~\eqref{eq-NbodyOT} for general pair costs of form $c_{\rm ee}({\bf r}_1,\dots,{\bf r}_N) = \sum_{i<j} w({\bf r}_i-{\bf r}_j)$, or Eq.~\eqref{nabla-v} for fully general costs.
Using Theorem \ref{theo-derivative} and \ref{theo-potentials}, we can derive the Kohn-Sham equations corresponding to the SCE energy functional \eqref{eq-energy-sce} with a computable effective potential.
It is the Euler-Lagrange equation corresponding to this minimization problem
(after a unitary transformation to diagonalize the symmetric $N\times N$ matrix of Lagrange multipliers):
find $\lambda_i\in\mathbb{R}, ~\phi_i\in H^1(\mathbb{R}^3)~(i=1,2,\dots,N)$ such that
\begin{equation}\label{ks-sce}
\left\{ \begin{array}{rcl}
\Big(-\frac{1}{2}\Delta + v_{\rm ext} + v_{\rm SCE}[\rho_{\Phi}]\Big)
\phi_i &=& \lambda_i\,\phi_i\quad \text{in}\quad\mathbb{R}^3, \quad i=1,2,\dots, N, \\[1ex]
\displaystyle \int_{\mathbb{R}^3}\phi_i\,\phi_j &=& \delta_{ij}.
\end{array} \right.
\end{equation}
This is a nonlinear eigenvalue problem, where the potential $v_{\rm SCE}[\rho_{\Phi}]$ depends on the electron density $\rho_{\Phi}$ associated with the orbitals $\phi_i$.
A self-consistent field (SCF) iteration algorithm is commonly resorted to for this nonlinear problem.
In each iteration step of the algorithm, a new effective potential is constructed from a trial electron
density and a linear eigenvalue problem is then solved to obtain the low-lying eigenvalues.
We shall comment further on the additive constant in \eqref{sce-kantorovich}. The above equations remain valid when $v_{\rm SCE}[\rho_{\Phi}]$ is modified by an arbitrary additive constant. This yields the same Kohn-Sham orbitals $\phi_i$ and only leads to a corresponding shift of the nonlinear eigenvalues $\lambda_i$. However, as pointed out in \cite{friesecke13}, it is only when $v_{\rm SCE}[\rho_{\Phi}]$ is precisely the Kantorovich potential that the ground state energy can equal the sum of Kohn-Sham eigenvalues, i.e.,
\begin{equation*}
E_0 = \sum_{i=1}^N \lambda_i.
\end{equation*}
In summary, an SCF algorithm for solving the Kohn-Sham equation \eqref{ks-sce} is given by
\begin{algorithm}\label{algorithm-ks_SCE}
SCF iterations for SCE-based Kohn-Sham equations
\begin{enumerate}
\item Given $\epsilon>0$. Let $k=0$ and $\rho_0$ be an initial electron density.
\item Calculate the co-motion functions from $\rho_k$
(by using the numerical methods introduced in the next section).
\item Calculate the effective potential $v_{\rm SCE}[\rho_k]$ by \eqref{eq-NbodyOT}.
\item Solve the linear eigenvalue problem
$$
\left(-\frac{1}{2}\Delta + v_{\rm ext} + v_{\rm SCE}[\rho_{\Phi}]\right)\phi_i
= \lambda_i\phi_i \quad i=1,\dots,N
$$
for low-lying eigenvalues to obtain new Kohn-Sham orbitals,
from which a new electron density $\rho_k^{\rm out}$ can be calculated.
\item If $\|\rho_k-\rho_k^{\rm out}\|<\epsilon$, stop; else,
generate a new electron density $\rho_{k+1}$ by some charge mixing technique
and go to 2.
\end{enumerate}
\end{algorithm}
\section{Numerical discretizations of optimal \\ transportation}
\label{section-numerical}
In each iteration of the SCF algorithm for solving \eqref{ks-sce}, one has to construct $v_{\rm SCE}[\tilde{\rho}]$ from a trial electron density $\tilde{\rho}$. According to \eqref{eq-NbodyOT}, this requires the solution of the optimal transport problem \eqref{energy-sce} with a given single-particle density to obtain the co-motion functions $T_i,~i=2,\dots,N$.
For simplicity, we only consider the case $N=2$, where only one co-motion function has to be calculated (which is denoted by $T$ in the following). For systems with more than two electrons, we refer to Section \ref{section-perspective} for a future perspective.
We discretize the computational domain into $n$ finite elements $e_1,\dots,e_n$. (We replace $\mathbb{R}^3$ by a bounded domain so that it can be discretized into a finite number of elements. This is reasonable since the electron density $\rho({\bf r})$ of a confined system decays exponentially fast to zero as $|{\bf r}|\rightarrow\infty$ \cite{hoffmann01}.) Each element is represented by a point ${\bf a}_k$ located at its barycenter and its electron mass
$\rho_k=\int_{e_k}\rho({\bf r})d{\bf r}$. Within this discretization, we can approximate the two-particle density $|\Psi({\bf r}_1,{\bf r}_2)|^2$ by a matrix $X=(x_{kl}) \in \mathbb{R}^{n \times n}$ with $x_{kl}=|\Psi({\bf a}_k,{\bf a}_l)|^2$. (Alternatively, one could identify the entries with the average $x_{kl} = \frac{1}{|e_k| \cdot |e_l|}\int_{e_k}\int_{e_l} |\Psi({\bf r}_1,{\bf r}_2)|^2 d{\bf r}_1 d{\bf r}_2$.)
The continuous problem \eqref{energy-sce} is then discretized into
\begin{equation}\label{ot_linprog}
\begin{array}{rl}
\displaystyle \min_{X} & \displaystyle \sum_{1\leq k,l\leq n} \frac{x_{kl}}{|{\bf a}_k-{\bf a}_l|} \\ \\
{\rm s.t.} & \sum_{1\leq k\leq n}x_{kl} = \frac{1}{2}\rho_k, \quad l=1,\dots,n \\[1ex]
& \sum_{1\leq l\leq n}x_{kl}=\frac{1}{2}\rho_l, \quad k=1,\dots,n \\[1ex]
& x_{kl}\geq 0. \end{array}
\end{equation}
Note that \eqref{ot_linprog} is a linear programming problem of the form
\begin{equation*}
\begin{split}
&\min_x f^T x \\
&{\rm s.t.}\quad A x = b ~{\rm and}~ x_k \ge 0,
\end{split}
\end{equation*}
where $x$ is the vector containing the entries of $X$. We can solve this problem by standard optimization routines like `\emph{linprog}' in \textsf{Matlab}. Due to the symmetry of the problem, one can assume that $x_{lk} = x_{kl}$ and only needs to consider $x_{kl}$ for $k \le l$.
As a remark, the dual problem of \eqref{ot_linprog} (in the sense of linear programming) results in a discretized version of the Kantorovich dual formulation \eqref{kantorovich}.
The solution of \eqref{ot_linprog} entails an approximation of the co-motion functions at the barycenters $\{{\bf a}_k\}_{1\leq k\leq n}$ via the matrix $X=(x_{kl})$:
\begin{equation}\label{T_approximate}
T_n({\bf a}_k)=\sum_{l=1}^n {\bf a}_l\frac{2 x_{kl}}{\rho_l},
\quad k=1,\dots,n,
\end{equation}
where $x_{kl}$ can also be regarded as the mass of electron transported from ${\bf a}_k$ to ${\bf a}_l$.
If the discretization is sufficiently fine, i.e., $n$ large enough, then $T_n$ is a good
approximation of $T$ (see the following numerical example).
For a \emph{uniform} discretization $\{e_k\}_{1\leq k\leq n}$,
the degrees of freedom for linear programming \eqref{ot_linprog} may be huge.
To reduce the computational cost, we use a locally refined mesh instead, which has more elements where the electron
density is high and less elements where the electron density is low.
Generally speaking, the optimal mesh may be such that each element $e_k$ has almost equal electron mass $\rho_k$.
This type of mesh can be generated by an adaptive procedure, say, one refines the element when its electron mass is
larger than a given threshold and coarse it otherwise.
Since the electron density decays exponentially fast to zero as $|{\bf r}|\rightarrow\infty$, the mesh is much coarser
far away from the nuclei than close to the nuclei, which reduces the degrees of freedom significantly.
As a remark, let us assume for a moment that all elements have exactly the same mass, $\rho_k = \bar{\rho}$ for all $k$. Then the constraints in \eqref{ot_linprog} force $X$ to be a \emph{doubly stochastic} matrix (up to a global scaling factor) with nonnegative entries. According to Birkhoff's theorem, the extremal points of the convex set of admissible matrices $X$ are the permutations, i.e., matrices with exactly one non-zero entry $\frac{1}{2} \bar{\rho}$ in each row (or column). Since the optimum is obtained at an extremal point, the optimizer can be chosen of this form. We have thus derived a discrete analogue of the Monge formulation, since the sum on the right of Eq.~\eqref{T_approximate} will have exactly one nonzero term.
Another important technique to reduce the computational cost is to exploit the symmetry of the system.
If the electron density $\rho$ has some kind of symmetric property, then we can reduce the computations
to some subdomain accordingly.
For example, \cite{seidl07} gives an explicit formula of co-motion functions for spherically symmetric
electron densities by making use of the symmetry. More precisely, it is proven that if the density has the form
$\rho({\bf r})=h(|{\bf r}|)$ with some function $h: [0,\infty) \rightarrow \mathbb{R}$,
then the corresponding co-motion function $T$ has to be spherically symmetric itself, that is
\begin{equation*}
T({\bf r})=g(|{\bf r}|)\frac{\bf r}{|\bf r|},\quad\forall~{\bf r}\in\mathbb{R}^3
\end{equation*}
with some function $g: [0,\infty) \rightarrow \mathbb{R}$.
This reduces the three-dimensional spherically symmetric problem into a one-dimensional problem.
Here we consider cylindrically symmetric systems, for instances, biatomic molecules.
The following theorem states that the co-motion function $T$ inherits the cylindrical symmetriy of the density.
\begin{theorem}\label{theo-symmetric}
Let $N=2$ and denote the cylindrical coordinates by $(\gamma,\varphi,z)$.
If $\rho({\bf r})=\varrho(\gamma,z)$ with some function $\varrho:[0,\infty)\times\mathbb{R}\rightarrow\mathbb{R}$,
then the corresponding co-motion function $T$ satisfies
\begin{equation}\label{T_symmetric}
T:~(\gamma,\varphi,z)\mapsto(\gamma',\varphi+\pi,z') \quad\forall~(\gamma,z)\in [0,\infty)\times\mathbb{R},
\end{equation}
where $(\gamma',z')=\ell(\gamma,z)$ with some map
$\ell:[0,\infty)\times\mathbb{R}\rightarrow [0,\infty)\times\mathbb{R}$.
\end{theorem}
\begin{proof}
Let $\mathcal{R}_{\theta}$ be the rotation operator with angle $\theta$ around the z-axis, i.e., $\mathcal{R}_{\theta}(\gamma,\varphi,z)=(\gamma,\varphi+\theta,z)$.
Let $\rho_2$ be a minimizer of \eqref{energy-sce} with single-particle electron density $\rho$,
such that $\rho_2({\bf r}_1,{\bf r}_2)=\frac{\rho({\bf r}_1)}{2}\delta({\bf r}_2-T({\bf r}_1))$
with $T$ the corresponding co-motion function.
Let $\tilde{\rho}_2({\bf r}_1,{\bf r}_2) = \rho_2(\mathcal{R}_{\theta}{\bf r}_1,\mathcal{R}_{\theta}{\bf r}_2)$.
We claim that $\tilde{\rho}_2$ is also a minimizer of \eqref{energy-sce}.
To see this, we observe that
\begin{equation*}
\tilde{\rho}_2({\bf r}_1,{\bf r}_2) = \frac{\rho(\mathcal{R}_{\theta}{\bf r}_1)}{2}
\delta(\mathcal{R}_{\theta}{\bf r}_2-T(\mathcal{R}_{\theta}{\bf r}_1)),
\end{equation*}
which satisfies the marginal constraint
\begin{equation}\label{proof-6-a}
\tilde{\rho}_2\mapsto\rho
\end{equation}
since $\rho({\bf r})=\rho(\mathcal{R}_{\theta}{\bf r})$.
Moreover, we have
\begin{equation}\label{proof-6-b}
\int_{\mathbb{R}^6}\frac{\tilde{\rho}_2({\bf r}_1,{\bf r}_2)}{|{\bf r}_1-{\bf r}_2|}d{\bf r}_1d{\bf r}_2
= \int_{\mathbb{R}^6}\frac{\rho_2({\bf r}_1,{\bf r}_2)}{|{\bf r}_1-{\bf r}_2|}d{\bf r}_1d{\bf r}_2
\end{equation}
from the fact that the cost function $\frac{1}{|{\bf r}_1-{\bf r}_2|}$ is invariant under the map
$({\bf r}_1,{\bf r}_2)\rightarrow(\mathcal{R}_{\theta}{\bf r}_1,\mathcal{R}_{\theta}{\bf r}_2)$.
\eqref{proof-6-a} and \eqref{proof-6-b} together imply that $\tilde{\rho}_2$ is a minimizer of \eqref{energy-sce}.
Therefore, $\tilde{T}:{\bf r}\mapsto\mathcal{R}_{\theta}^{-1}T(\mathcal{R}_{\theta}{\bf r})$
is also a co-motion function of this problem.
Because the co-motion function is unique in the case of two particles (see \cite{cotar13a}), we have
\begin{equation}\label{proof-3-a}
T(\mathcal{R}_{\theta}{{\bf r}})=\mathcal{R}_{\theta}T({\bf r}).
\end{equation}
Since we minimize
\begin{equation*}
\begin{split}
\int_{\mathbb{R}^3}\frac{\rho({\bf r})}{|{\bf r}-T({\bf r})|}d{\bf r}
&= \int_0^{\infty}\gamma d\gamma\int_{-\infty}^{\infty}dz \int_0^{2\pi}d\varphi
\frac{\rho((\gamma,\varphi,z))}{|(\gamma,\varphi,z)-T((\gamma,\varphi,z))|}d\varphi \\
&\stackrel{{\rm fix}~\varphi}{=} \int_0^{\infty}\gamma d\gamma\int_{-\infty}^{\infty}dz
\int_0^{2\pi}d\theta \frac{\rho(\mathcal{R}_{\theta}(\gamma,\varphi,z))}{|\mathcal{R}_{\theta}
(\gamma,\varphi,z) - T(\mathcal{R}_{\theta}(\gamma,\varphi,z))|}d\theta \\
&\stackrel{\eqref{proof-3-a}}{=} \int_0^{\infty}\gamma d\gamma\int_{-\infty}^{\infty}dz
\int_0^{2\pi}d\theta \frac{\rho(\mathcal{R}_{\theta}(\gamma,\varphi,z))}{|\mathcal{R}_{\theta}
(\gamma,\varphi,z) - \mathcal{R}_{\theta}T((\gamma,\varphi,z))|}d\theta \\
&= 2\pi \int_0^{\infty} \int_{-\infty}^{\infty}
\frac{\gamma\varrho((\gamma,z))}{|(\gamma,\varphi,z) - T((\gamma,\varphi,z))|}d\gamma dz,
\end{split}
\end{equation*}
$T$ has to satisfy \eqref{T_symmetric} to maximize the denominator.
This completes the proof.
\end{proof}
{\bf Example.} We take a one-dimensional two-electron system as an example to illustrate our numerical method.
Although this type of algorithm is not particularly interesting for one-dimensional systems
since analytical formulations are known \cite{malet14,friesecke13},
it is more suitable to present the numerical results and explain our idea.
Let $\Omega=[-5,5]$ and $\rho(x)=0.4-0.08|x|$, see Figure \ref{fig-rho1d}.
The exact co-motion function can be calculated explicitly
\begin{equation}\label{exact1d_comotion}
T(x)=\begin{cases}
\hspace{8pt} 5\left(1-\big[1-0.5(x+5)(0.4+0.08x)\big]^{1/2}\right) & {\rm if}~x\leq 0, \\
-5\left(1-\big[1-0.5(-x+5)(0.4-0.08x)\big]^{1/2}\right) & {\rm if}~x>0.
\end{cases}
\end{equation}
We observe in Figure \ref{fig-comotion} that the numerical approximations of the co-motion function
can be very accurate (compared with the exact formula \eqref{exact1d_comotion}).
Figure~\ref{fig-convergence} shows the convergence of the uniform numerical approximations.
The nonuniform mesh (red elements in Figure~\ref{fig-rho1d}) achieves a much higher accuracy with the same degrees of freedom, see Figure~\ref{fig-adaptive}.
Actually, we observe that the errors of a nonuniform mesh with $n=20$ is even smaller compared to a uniform mesh with $n=40$ on average.
\begin{figure}[!ht]
\centering
\subfloat[electron density]{
\label{fig-rho1d}
\includegraphics[width=0.45\textwidth]{rho1d_mesh}} \quad
\subfloat[co-motion function]{
\label{fig-comotion}
\includegraphics[width=0.45\textwidth]{co-motion}} \\
\subfloat[$T_n$ error, uniform mesh]{
\label{fig-convergence}
\includegraphics[width=0.45\textwidth]{Tn_error}} \quad
\subfloat[$T_n$ error, nonuniform mesh]{
\label{fig-adaptive}
\includegraphics[width=0.45\textwidth]{adaptive_error}}
\caption{(a) The one-dimensional electron density $\rho$ and a nonuniform discretization.
(b) The co-motion function corresponding to $\rho$ and its approximation.
(c) Numerical errors of $T_n$ using uniform meshes, and (d) a nonuniform mesh.}
\end{figure}
Remember that the aforementioned electron density is symmetric in the sense of
\begin{equation*}
\rho(x)=\rho(-x).
\end{equation*}
Therefore, the corresponding co-motion function is symmetric itself, i.e.
\begin{equation*}
T(x) = -T(-x),
\end{equation*}
which maps $[-5,0]$ to $[0,5]$ and maps $[0,5]$ to $[-5,0]$.
Hence, it is only necessary to calculate $T(x)$ on half of the domain, say, $[-5,0]$.
An important application of the numerical methods introduced above is to simulate the H$_2$
molecule at its dissociating limit. We provide more details in the next section.
\section{H$_2$ bond disassociation}
\label{section-h2}
We consider H$_2$ molecule in this section.
Let $R>0$ and ${\bf R}_A=(-R,0,0),~{\bf R}_B=(R,0,0)$ be the locations of two hydrogen atoms. Physically, the hydrogen molecule should dissociate into two free hydrogen atoms as the bond length $2R\rightarrow\infty$, with the ground state spin-unpolarized.
The spin-restricted Hartree-Fock and Kohn-Sham DFT models give the correct spin multiplicity,
but overestimate total energies, i.e., higher than that of two free hydrogen atoms.
In comparison, the spin-unrestricted models give fairly good total energies, while the wave
functions are spin-contaminated, which is known as ``symmetry breaking'' in H$_2$ bond dissociation.
Here we focus on the SCE-DFT model without symmetry breaking, and
show both theoretically and numerically that the restricted Kohn-Sham model \eqref{min-sce}
gives the correct ground state energy in the dissociation limit $R\rightarrow\infty$.
Denote by $e_0$ the ground state energy of a single hydrogen atom
\begin{equation}\label{e_0}
e_0=\inf\left\{\frac{1}{2}\int_{\mathbb{R}^3}|\nabla\phi({\bf r})|^2d{\bf r}
-\int_{\mathbb{R}^3}\frac{|\phi({\bf r})|^2}{|\bf r|}d{\bf r},
~\phi\in H^1(\mathbb{R}^3),~\|\phi\|_{L^2(\mathbb{R}^3)}=1\right\}.
\end{equation}
$E_{\rm SCE}(R)$ denotes the ground state energy of the hydrogen molecule in the SCE-DFT model \eqref{eq-energy-sce}
\begin{multline}\label{E_R}
E_{\rm SCE}(R) = \frac{1}{2R} + \inf \Big\{\int_{\mathbb{R}^3}|\nabla\phi({\bf r})|^2d{\bf r}
+ 2\int_{\mathbb{R}^3}v_{\rm ext}({\bf r})|\phi({\bf r})|^2d{\bf r}
+ V_{\rm ee}^{\rm SCE}[2|\phi|^2], \\
\phi\in H^1(\mathbb{R}^3),~\|\phi\|_{L^2(\mathbb{R}^3)}=1\Big\}, \quad
\end{multline}
where $v_{\rm ext}({\bf r})=-\frac{1}{|{\bf r}-{\bf R}_A|}-\frac{1}{|{\bf r}-{\bf R}_B|}$.
The following result indicates that the SCE-DFT model is correct for the H$_2$ molecule at its dissociating limit.
\begin{theorem}\label{theo-sce-h2}
Let $e_0$ and $E_{\rm SCE}(R)$ be given by \eqref{e_0} and \eqref{E_R} respectively.
We have
\begin{equation}\label{sce-h2-limit}
\lim_{R\rightarrow\infty}E_{\rm SCE}(R) = 2e_0.
\end{equation}
\end{theorem}
\begin{proof}
First, we establish an upper bound of $E_{\rm SCE}(R)$.
Let $\varphi(r)=e^{-r}/\sqrt{\pi}$ and
\begin{equation*}
\psi({\bf r}) = \left(\frac{1}{2} \Big(\varphi^2(|{\bf r}-{\bf R}_A|) + \varphi^2(|{\bf r}-{\bf R}_B|) \Big) \right)^{1/2}.
\end{equation*}
Note that $\varphi$ is the minimizer of \eqref{e_0},
and $\|\varphi\|_{L^2(\mathbb{R}^3)}=1$ implies $\|\psi\|_{L^2(\mathbb{R}^3)}=1$.
We have
\begin{equation}\label{proof-4-b}
E_{\rm SCE}(R) \leq \frac{1}{2R} + \int_{\mathbb{R}^3}|\nabla\psi({\bf r})|^2d{\bf r}
+ 2\int_{\mathbb{R}^3}v_{\rm ext}({\bf r})|\psi({\bf r})|^2d{\bf r} + V_{\rm ee}^{\rm SCE}[2|\psi|^2].
\end{equation}
Let $\phi_1({\bf r})=\varphi(|{\bf r}-{\bf R}_A|)$ and $\phi_2({\bf r})=\varphi(|{\bf r}-{\bf R}_B|)$.
A direct calculation leads to
\begin{equation}\label{proof-4-c}
\begin{split}
\int_{\mathbb{R}^3}|\nabla\psi|^2
&= \int_{\mathbb{R}^3}\frac{|\phi_1\nabla\phi_1+\phi_2\nabla\phi_2|^2}{4(\phi_1^2+\phi_2^2)} \\
&\leq \frac{1}{2} \int_{\mathbb{R}^3}(|\nabla\phi_1|^2+|\nabla\phi_2|^2) = \int_{\mathbb{R}^3}|\nabla\varphi|^2
\end{split}
\end{equation}
and
\begin{equation}\label{proof-4-d}
\begin{split}
& 2\int_{\mathbb{R}^3}v_{\rm ext}({\bf r})|\psi({\bf r})|^2d{\bf r}
= -\int_{\mathbb{R}^3}\frac{\phi_1^2({\bf r})+\phi_2^2({\bf r})}{|{\bf r}-{\bf R}_A|}d{\bf r}
-\int_{\mathbb{R}^3}\frac{\phi_1^2({\bf r})+\phi_2^2({\bf r})}{|{\bf r}-{\bf R}_B|}d{\bf r} \\
&= -2\int_{\mathbb{R}^3}\frac{\varphi^2(|{\bf r}|)}{|{\bf r}|}d{\bf r}
-\int_{\mathbb{R}^3}\frac{\phi_2^2({\bf r})}{|{\bf r}-{\bf R}_A|}d{\bf r}
-\int_{\mathbb{R}^3}\frac{\phi_1^2({\bf r})}{|{\bf r}-{\bf R}_B|}d{\bf r} \\
&\leq -2\int_{\mathbb{R}^3}\frac{\varphi^2(|{\bf r}|)}{|{\bf r}|}d{\bf r}.
\end{split}
\end{equation}
Let $\rho_2({\bf r}_1,{\bf r}_2)=2|\psi({\bf r}_1)|^2\delta({\bf r}_2,{\bf r}_1-{\bf R}_A+{\bf R}_B)$,
we have
\begin{equation}\label{proof-4-e}
V_{\rm ee}^{\rm SCE}[2|\psi|^2] \leq
\int_{\mathbb{R}^6}\frac{\rho_2({\bf r}_1,{\bf r}_2)}{|{\bf r}_1-{\bf r}_2|}d{\bf r}_1d{\bf r}_2
= \int_{\mathbb{R}^3}\frac{2|\psi({\bf r}_1)|^2}{|{\bf R}_A-{\bf R}_B|}d{\bf r}_1 = \frac{1}{R}.
\end{equation}
Taking \eqref{proof-4-b}, \eqref{proof-4-c}, \eqref{proof-4-d} and \eqref{proof-4-e} into account, we have
\begin{equation}\label{proof-4-f}
E_{\rm SCE}(R) \leq 2e_0 + \frac{3}{2R}.
\end{equation}
To give a lower bound of $E_{\rm SCE}(R)$, we observe that
\begin{equation}\label{proof-4-g}
\begin{split}
E_{\rm SCE}(R) &\geq \frac{1}{2R} + \inf_{\|\phi\|_{L^2}=1}
\left\{\int_{\mathbb{R}^3}|\nabla\phi({\bf r})|^2d{\bf r}
+ 2\int_{\mathbb{R}^3}v_{\rm ext}({\bf r})|\phi({\bf r})|^2d{\bf r} \right\} \\
&= \frac{1}{2R} + 2\inf_{\|\phi\|_{L^2}=1}\langle\phi | \hat{h} | \phi\rangle,
\end{split}
\end{equation}
where $\hat{h}$ is the H$_2^+$ Hamiltonian $\hat{h}=-\frac{1}{2}\Delta + v_{\rm ext}$.
We claim that
\begin{equation}\label{proof-4-j}
\langle\phi |\hat{h} | \phi\rangle \geq (e_0-O(R^{-1}))\|\phi\|^2_{L^2}
\quad \forall~\phi\in H^1(\mathbb{R}^3).
\end{equation}
To show \eqref{proof-4-j}, we first decompose the unity function on $\mathbb{R}$ into
two smooth cutoff functions $\tilde{\zeta}_1$ and $\tilde{\zeta}_2$,
such that $\tilde{\zeta}_1^2+\tilde{\zeta}_2^2=1$,
$\tilde{\zeta}_1(x)=0$ for $x<-\frac{1}{2}$, $\tilde{\zeta}_2(x)=0$ for $x>\frac{1}{2}$,
and $|\nabla\tilde{\zeta}_i|\leq C^*$ with some constant $C^*$.
Let
$$
\zeta_i({\bf r})=\zeta_i(x,y,z)=\tilde{\zeta}_i(x/R) \quad i=1,2
$$
and $\phi_i=\zeta_i\phi$. We have $|\nabla\zeta_i|\leq C^*/R$ and
\begin{equation*}
\sum_{i=1}^2|\nabla\phi_i|^2 = \sum_{i=1}^2 (|\nabla\zeta_i|^2)|\phi|^2 + |\nabla\phi|^2.
\end{equation*}
Therefore,
\begin{equation*}
\begin{split}
\langle\phi |\hat{h} | \phi\rangle &=
\frac{1}{2}\sum_{i=1}^2\int_{\mathbb{R}^3}|\nabla\phi_i|^2 - \int_{\mathbb{R}^3}\sum_{i=1}^2 (|\nabla\zeta_i|^2) |\phi|^2
+ \int_{\mathbb{R}^3} v_{\rm ext}\big(|\phi_1|^2+|\phi_2|^2\big) \\
&\geq \sum_{i=1}^2\int_{\mathbb{R}^3}\left(\frac{1}{2}|\nabla\phi_i|^2 - \frac{|\phi_i|^2}{|{\bf r}|}\right)
- \left(\frac{2C^*}{R}\right)^2 - \frac{2}{R},
\end{split}
\end{equation*}
which implies \eqref{proof-4-j}.
Therefore, we obtain from \eqref{proof-4-g} and \eqref{proof-4-j} that
\begin{equation}\label{proof-4-h}
E_{\rm SCE}(R) \geq 2e_0 + O(R^{-1}).
\end{equation}
Together with the upper bound \eqref{proof-4-f}, this leads to \eqref{sce-h2-limit} at the limit $R\rightarrow\infty$.
\end{proof}
\begin{figure}[!ht]
\centering
\subfloat{\includegraphics[width=0.5\textwidth]{h2_rhoz}}
\subfloat{\includegraphics[width=0.5\textwidth]{h2_meshz}}
\caption{The electron density and corresponding mesh on slice $z=0$ ($R = 5$).}
\label{fig_h2_rho_mesh}
\end{figure}
In what follows, we present a numerical simulation of the dissociating H$_2$ molecule to support the theory.
The computations are carried out on a bounded domain $\Omega=[-10,10]^3$.
We use Algorithm~\ref{algorithm-ks_SCE} to solve \eqref{ks-sce} for the ground state energies
and electron densities for different bond length $2R$.
Concerning the optimal transport problem, we use the numerical methods introduced in
Section \ref{section-numerical} and calculate the co-motion function in each SCF iteration step.
The nonuniform mesh is generated by the package \textsf{PHG} \cite{phg},
a toolbox for parallel adaptive finite element programs developed at the State Key Laboratory of
Scientific and Engineering Computing of the Chinese Academy of Sciences. While the problem is effectively two-dimensional according to Theorem~\ref{theo-symmetric}, we have performed the calculations in three dimensions since \textsf{PHG} is tailored to three-dimensional problems. Figure~\ref{fig_h2_rho_mesh} shows a contour plot of
the electron density at slice $z=0$ and the corresponding mesh.
One observes that the grid reflects the higher density around the nuclei.
\begin{figure}[!ht]
\centering
\includegraphics[height=5.5cm]{h2_mesh1}\hskip 3.0cm
\includegraphics[height=5.5cm]{h2_mesh3}
\put(-140,70){\makebox(1,1){$x\geq 0,y\geq 0,z\geq 0,z\leq y$}}
\put(-120,60){\makebox(1,1){$\Longrightarrow$}}
\caption{Symmetric decomposition of the computation domain $\Omega$ for H$_2$.}
\label{fig_h2_mesh}
\end{figure}
To further reduce the computational cost, we can exploit the cylindrical symmetry of the system with the help of Theorem~\ref{theo-symmetric}. As shown in Figure~\ref{fig_h2_mesh}, the degrees of freedom can be reduced to $1/16$ of the original volume.
For the linear programming problem \eqref{ot_linprog}, we resort to \textsf{MOSEK} \cite{mosek}, a high-performance software for large-scale optimization problems.
The computational results are presented in Figure~\ref{fig_h2_bond}, in which we compare the bond energies in dependence of $R$ using the LDA and SCE Kohn-Sham methods, respectively. Here, the bond energies are the ground state energies of the systems minus $2e_0$, which is expected to be zero when the two hydrogen atoms are disassociated. Note that the SCE model shows the correct asymptotic behavior, while the LDA model fails at large $R$ by giving too large energies. For comparison, the LDA error $0.065\,{\rm a.u.}$ in \cite{ChallengesDFT2012} for the infinitely stretched H$_2$ molecule is lower than in Figure~\ref{fig_h2_bond} since twice the LDA-hydrogen energy is subtracted instead of twice the exact $e_0$, but remains significant; errors of similar magnitude are reported there for other functionals such as B3LYP or PBE.
\begin{figure}[!ht]
\centering
\includegraphics[width=0.7\textwidth]{h2-bond}
\caption{H$_2$ potential energy curve as a function of the bond length for the SCE and LDA models.}
\label{fig_h2_bond}
\end{figure}
As physical explanations for these results, at long internuclear separations, if one electron is located near atom A, the other will be found close to atom B. This correlation is correctly reflected by the optimal transport model, hence the SCE model gives asymptotically the product of hydrogen orbitals on the two nuclei. In contrast, within the Kohn-Sham LDA framework, the two electrons are constrained to be in the same spatial orbital and each electron experiences only the average effect of the other, thus each electron has equal probability of being near A or B, irrespective of the position of the other electron. The possibility of both electrons being on the same atom is not excluded, as reflected in the wrong asymptotic behavior of the disassociation energy in Figure~\ref{fig_h2_bond}.
\begin{figure}[!htb]
\centering
\subfloat[]{\includegraphics[width=0.5\textwidth]{h2_rho_x}}
\subfloat[]{\includegraphics[width=0.5\textwidth]{h2_vcu_x}}
\caption{The SCE and LDA electron densities (a) and potentials (b) of H$_2$, plotted along the molecular axis.}
\label{fig_h2_x_rho_vcu}
\end{figure}
We further compare the electron densities and Coulomb potentials obtained by the two different models in Figure~\ref{fig_h2_x_rho_vcu}. Note that the scalar offset of the potential in the figure is determined by the boundary conditions, i.e., $\frac{N}{R}$ for LDA and $\frac{N-1}{R}$ for SCE \cite{seidl07}. The electron densities are actually quite close, while the shape of the potentials differs substantially when $R$ is large. The SCE potential is larger between the hydrogen atoms, favoring a depletion of the bond charge whenever the two atoms separate. This produces the correct results for large $R$.
\begin{figure}[!htb]
\centering
\includegraphics[width=0.4\textwidth]{map_contour}
\caption{Optimal transport mapping of the region $0.04 \le \rho({\bf r}) \le 0.08$ (indicated by the blue dots inside the red contours) to the green area, for the H$_2$ molecule with $R = 1$. Each blue dot corresponds to one barycenter in the numerical discretization, and has been rotated into the $x$-$y$-plane with $y \ge 0$ for visual clarity. The green dots are precisely the images of the blue dots under the optimal transport map.}
\label{fig_h2_ot_map}
\end{figure}
Finally, as illustration of the co-motion function or optimal map, Figure~\ref{fig_h2_ot_map} shows the image of the map on a density contour.
\section{Conclusions and perspectives}
\label{section-perspective}
The numerical discretization of the SCE optimal transport problem with Coulomb cost and two marginals leads to the linear programming problem \eqref{ot_linprog}, which can indeed be solved in practice, as we have demonstrated in a proof of concept calculation. The self-consistent SCE-DFT simulation of the dissociating H$_2$ molecule agrees well with the physically correct limit, unlike standard DFT models like LDA.
The theory of this paper applies to more general systems with arbitrary numbers of electrons,
however, the numerical algorithms need further developments. Specifically, we can restrict the $N$-particle density to the ansatz
\begin{equation*}
\rho_N({\bf r}_1,\dots,{\bf r}_N) = \frac{\rho({\bf r}_1)}{N}
\gamma_2({\bf r}_1,{\bf r}_2)\gamma_3({\bf r}_1,{\bf r}_3)
\cdots\gamma_N({\bf r}_1,\gamma_N),
\end{equation*}
where $\rho$ is the given single-particle density.
Here, $\gamma_j({\bf r}_1,{\bf r}_j)$ represents the probability of the $j$th electron being found at ${\bf r}_j$
while the first electron is located at ${\bf r}_1$.
We have
\begin{equation*}
\int_{\mathbb{R}^3}\gamma_j({\bf r}_1,{\bf r}_j)d{\bf r}_j = 1 \quad\quad j=2,\dots,N.
\end{equation*}
With a given discretization $\{e_k\}_{1\leq k\leq n}$ and barycenters ${\bf a}_k$ of $e_k$,
we can approximate $\gamma_j({\bf r}_1,{\bf r}_j)$ by a matrix $X_j=(x_{j,kl}) \in \mathbb{R}^{n\times n}$ for $j=2,\dots,N$.
Then the continuous model \eqref{energy-sce} is reduced to
\begin{equation}\label{ot_quadprog}
\begin{array}{rl}
\displaystyle \min_{X_2,\dots,X_N} &
\displaystyle \sum_{1 < j \leq N} \sum_{k,l=1}^n \frac{x_{j,kl}}{|{\bf a}_k - {\bf a}_l|}
\cdot \frac{\rho_k}{N}
+ \sum_{1 < i < j\leq N} \sum_{k,l,l' = 1}^n \frac{x_{i,kl} \cdot x_{j,kl'}}{|{\bf a}_l - {\bf a}_{l'}|}
\cdot \frac{\rho_k}{N} \\ \\
\mathrm{s.t.} & \sum_{k=1}^n x_{i,kl}=1, \quad l = 1, \dots, n, \quad i = 2, \dots, N \\[1ex]
& \sum_{l=1}^n x_{i,kl} = 1, \quad k = 1, \dots, n, \quad i = 2, \dots, N \\[1ex]
& x_{i,kl} \ge 0.
\end{array}
\end{equation}
This is a quadratic programming problem of the form
\begin{align*}
&\min_x \, x^T H x + f^T x \\
&\mathrm{s.t.}\quad A\,x = b \quad\mathrm{and}\quad x_k \ge 0.
\end{align*}
By solving the above quadratic programming problem, we approximate the co-motion functions
$T_2,\dots,T_N$ by the matrices $X_2=(x_{2,kl}), \dots, X_N = (x_{N,kl})$.
Similar to \eqref{T_approximate}, the co-motion functions can be approximated by
$$
T_i({\bf a}_k) \approx \sum_{1\leq l\leq n} {\bf a}_l \, x_{i,kl}
\qquad k=1,\dots,n, \quad i=2,\dots,N.
$$
However, a serious difficulty in solving \eqref{ot_quadprog} stems from the non-convexity of the matrix $H$. Moreover, the symmetric decomposition is not clear for systems with more than two electrons. We plan to investigate these issues in future work.
\medskip
{\bf Acknowledgments}\quad C.M.\ acknowledges support from the DFG project FR~1275/3-1.
|
\section{The main result}\label{S:Main}
The \text{\sc Meat-axe}\ is an algorithm often used to test whether a given group
or algebra of matrices over a finite field acts irreducibly on the
underlying vector space, see \cite{P,HR,NP2}. It uses random selection
to find a `good' matrix, and if successful is able to determine
whether the action is reducible or irreducible. One definition of a
`good' matrix in this context is a {\it cyclic} matrix. (A matrix is
cyclic if its characteristic and minimal polynomials are
equal.) The density of cyclic matrices in absolutely irreducible
groups and algebras is constrained by the following result of Neumann
and the fourth author \cite[Theorem~4.1]{NP}. The probability
$P_{d,q}:=\textup{Prob}(\textup{$X\in\textup{M}(d,\mathbb{F}_q)$ is non-cyclic})$ satisfies
\begin{equation}\label{E:NP}
\frac{q^{-3}}{1+q^{-1}}<P_{d,q}<\frac{q^{-3}}{(1-q^{-1})(1-q^{-2})}
\qquad\textup{for all $d\geq2$ and $q\geq2$}.
\end{equation}
Thus $\frac{2q^{-3}}{3}\leq P_{d,q}\leq\frac{8q^{-3}}{3}$, so
$P_{d,q}=\Omega(q^{-3})$ for $d\geq2$. If $d=1$, then $P_{1,q}=0$ because each
$1\times1$ matrix is cyclic. Bounds on the proportion of non-cyclic matrices in
irreducible-but-not-absolutely-irreducible matrix algebras are also available
in \cite{NP}.
This note shows that cyclic matrices are less dense in maximal
{\it reducible} matrix algebras than full matrix algebras, with density
$1-c(q)q^{-2}$ rather than $1-c'(q)q^{-3}$ where $c(q),c'(q)$ are bounded
functions. We do not know how to estimate the density $\delta$ of cyclic
matrices in arbitrary non-maximal reducible algebras.
Since $0\leq\delta\leq1$, our lower bound
$q^{-2}\left(1+c_1q^{-1}\right)<\delta$ is unhelpful if $c_1<-q$ for some
choice of~$q$. Similarly, our upper bound
$\delta<q^{-2}\left(1+c_2q^{-1}\right)$ is unhelpful if
$c_2>q(q^2-1)$. We go to some effort to find helpful bounds
for all values of $q$. While motivated by a complexity analysis of the
\text{\sc Meat-axe}\ algorithm, we feel that this problem has broader interest.
A modification of Norton's Irreducibility Test, called the Cyclic
Irreducibility Test, was presented in \cite{NP2}. It was shown to be
a Monte Carlo algorithm that proved irreducibility of a finite
irreducible matrix algebra $\A$ provided a \emph{cyclic pair}
was found, that is a pair $(v,X)$ where $X$ is a cyclic matrix in $\A$,
and $v$ is a cyclic vector for $X$. It was hoped that cyclic
pairs in reducible matrix algebras, if such exist, could be used to
construct a proper $\A$-invariant subspace. However, it
was not known which reducible algebras $\A$ might contain a
sufficiently high proportion of cyclic matrices to make this approach
worth exploring. In this paper we prove that finite maximal
reducible matrix algebras do indeed have a
plentiful supply of cyclic elements, with the proportion slightly less
than that for the full matrix algebra. A variant of the Cyclic
Irreducibility Test is given in~\cite[p.\;141]{B}.
\goodbreak
\noindent
{\bf Notation A.}\quad
The following notation will be used throughout the paper.\par
\hi{$F=\mathbb{F}_q$ a finite field with $q$ elements;}
\hi{$V=F^n$ the $F$-space of $1\times n$ row vectors;}
\hi{$U$ a fixed $r$-dimensional subspace of $V$ where $0<r<n$;}
\hi{$\textup{M}(V)=\textup{M}(n,F)=F^{n\times n}$ the $F$-algebra of all $n\times n$ matrices
over $F$;}
\hi{$\GL(V)$ the group of units of $\textup{M}(V)$: isomorphic to the general linear
group $\GL(n,q)$;}
\hi{$\textup{M}(V)_U$ the stabilizer in $\textup{M}(V)$ of $U$: isomorphic to the algebra of matrices
$X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)$ with
$A\in F^{r\times r}$, $B\in F^{(n-r)\times(n-r)}$, and
$C\in F^{(n-r)\times r}$;}
\hi{$\GL(V)_U$ the group of units of $\textup{M}(V)_U$ comprising all $X$ with
$\det(X)\kern-1pt=\kern-1pt\det(A)\det(B)\neq0$.}
\setlength{\parskip}{5pt}
\begin{theorem}\label{T:Main}
Suppose that $0<r<n$ and $U$ is an $r$-dimensional subspace of $V:=\mathbb{F}_q^n$.
Then there exist constants $c_1,c_2$, independent of $n,r,q$, such that
the probability that a uniformly distributed random matrix
$X\in\textup{M}(V)_U$ is non-cyclic satisfies
\[
q^{-2}(1+c_1q^{-1})\leq\textup{Prob}(\textup{$X\in\textup{M}(V)_U$ is non-cyclic})
\leq q^{-2}(1+c_2q^{-1}).
\]
The constants $c_1=-\frac{4}{3}$ and $c_2=\frac{35}{3}$ suffice.
\end{theorem}
\begin{table}[!ht]
\[
\begin{array}{cc}
\hline
\dim U&\mbox{Proportion of cyclic matrices in $\textup{M}(V)_U$ as $\dim(V)\to\infty$}\\ \hline
1&1-q^{-2}-2q^{-3}-\hspace{0.17cm}q^{-4}\hspace{1.3cm}+\hspace{0.18cm}2q^{-6}
+\hspace{0.18cm}3q^{-7}\hspace{0.08cm} +\textup{lower terms}\\
2&1-q^{-2}-4q^{-3}-\hspace{0.17cm}q^{-4}+4q^{-5}+\hspace{0.18cm}5q^{-6}
+\hspace{0.18cm}4q^{-7}\hspace{0.05cm}+\textup{lower terms}\\
3&1-q^{-2}-4q^{-3}-3q^{-4}+4q^{-5}+11q^{-6}+\hspace{0.18cm}8q^{-7}
+\textup{lower terms}\\
4&1-q^{-2}-4q^{-3}-3q^{-4}+2q^{-5}+11q^{-6}+14q^{-7}+\textup{lower terms}\\
5&1-q^{-2}-4q^{-3}-3q^{-4}+2q^{-5}+\hspace{0.18cm}9q^{-6}+14q^{-7}
+\textup{lower terms}\\
6&1-q^{-2}-4q^{-3}-3q^{-4}+2q^{-5}+\hspace{0.18cm}9q^{-6}+12q^{-7}
+\textup{lower terms}\\
7&1-q^{-2}-4q^{-3}-3q^{-4}+2q^{-5}+\hspace{0.18cm}9q^{-6}+12q^{-7}
+\textup{lower terms}\\ \hline
\end{array}
\]
\caption{Proportions of cyclic matrices in $\textup{M}(V)_U$ as $\dim(V)\to\infty$.}\label{tm}
\end{table}
\begin{table}[!ht]
\[
\begin{array}{cc}
\hline
\dim U&\mbox{Proportion of cyclic matrices in $\GL(V)_U$ as $\dim(V)\to\infty$}\\ \hline
1&1 - q^{-2} - 2q^{-3} \hspace{1.1cm} + \hspace{0.18cm} q^{-5}
+ 3q^{-6 }\hspace{0.05cm} + q^{-7}\hspace{0.2cm} +\textup{lower terms}\\
2&1 - q^{-2} - 3q^{-3} + q^{-4} + 3q^{-5} + 4q^{-6}-2q^{-7} +\textup{lower terms}\\
3&1 - q^{-2} - 3q^{-3} + q^{-4} + 4q^{-5} + 4q^{-6}-5q^{-7} +\textup{lower terms}\\
4&1 - q^{-2} - 3q^{-3} + q^{-4} + 4q^{-5} + 4q^{-6}-6q^{-7} +\textup{lower terms}\\
5&1 - q^{-2} - 3q^{-3} + q^{-4} + 4q^{-5} + 4q^{-6}-6q^{-7} +\textup{lower terms}\\
6&1 - q^{-2} - 3q^{-3} + q^{-4} + 4q^{-5} + 4q^{-6}-6q^{-7} +\textup{lower terms}\\
7&1 - q^{-2} - 3q^{-3} + q^{-4} + 4q^{-5} + 4q^{-6}-6q^{-7} +\textup{lower terms}\\
\hline
\end{array}
\]
\caption{Proportions of cyclic matrices in $\GL(V)_U$ as $\dim(V)\to\infty$.}\label{tgl}
\end{table}
\begin{remark}\label{rem1} {\rm
(a) The lower bound in Theorem~\ref{T:Main} is positive for all
$q\geq2$, and the upper bound is less than~1 for all $q>2$. With
more care we may increase~$c_1$ and decrease~$c_2$. However, a new
argument is needed to give an upper bound less than~1 when $q=2$
because the first term in (\ref{E:lower}) below is
$\frac{q^{-2}}{(1-q^{-1})^2}=1$ when~$q=2$.
(b) The bounds in Theorem~\ref{T:Main} in the cases $r=1$ and $r=n-1$ can be
deduced from results in
Jason Fulman's paper \cite{F} since in these cases $\GL(V)_U$ is an affine
group. The first asymptotic estimate for the probability in
Theorem~\ref{T:Main}, for general values of $r$, was given as the
main result in the PhD thesis of the first author~\cite{B} where a
probabilistic generating function was found for the proportion of
cyclic matrices in $\GL(V)_U$ for a subspace $U$ of fixed
dimension $r$. The limiting proportions of cyclic matrices in both
$\GL(V)_U$ and $\textup{M}(V)_U$, as $\dim(V)\rightarrow\infty$, were
proved to be power series in $q^{-1}$ of the form $1-q^{-2}+
\sum_{i\geq3} \gamma_iq^{-i}$. (In Tables~\ref{tm} and~\ref{tgl} the `lower
terms' residual was not bounded by a function of $r$ and $q$ in~\cite{B}.
By contrast, bounding constants independent of $r$, $n$, $q$ are explicit
in the statement, and proof, of Theorem~\ref{T:Main}.) Exact values for
these limiting proportions can be determined from the generating
function for small values of~$r$, and some sample results are given
in Tables~\ref{tm} and~\ref{tgl}. These results show that the
$\gamma_i$ depend mildly on the dimension~$r$ when $i\geq3$. The
expressions for $r=1,2$ were deduced analytically, and those for
$3\leq r\leq 7$ were obtained using \textsc{Mathematica}~\cite{Mca}.
(c) Truncating the power series in Table~\ref{tm} suggests (heuristically) that
the probability in Theorem~\ref{T:Main} `ought' to have the form
$q^{-2}(1+4q^{-1})$. This is consistent with the constants
given in Theorem~\ref{T:Main} as
$-\frac{4}{3}=c_1\leq 4\leq c_2=\frac{35}{3}$.
(d) The PhD thesis of the first author contains analogous results for
the limiting proportions (as $\dim(V)\to\infty$) of cyclic
matrices in maximal completely reducible matrix
algebras~\cite[Theorems 5.2.8 and 5.3.5]{B}, see also the
unpublished paper~\cite{BGP}. The limiting proportions of separable matrices
in maximal reducible matrix algebras are described in~\cite[Theorem 6.4.6]{B}.
}
\end{remark}
\begin{proof}[Proof Strategy for Theorem~\textup{\ref{T:Main}}.]
Since $\GL(V)$ acts transitively on the set of $r$-dimensional subspaces
of $V$, the stabilizers of $r$-dimensional subspaces, being conjugate,
all have the same cardinality. Thus it suffices to consider the stabilizer
$\textup{M}(V)_U$ of the $r$-dimensional subspace
$U:=\langle e_1,\dots,e_r\rangle$ where
$e_i$ denotes the $i$th row of the $n\times n$ identity matrix $I_n$.
Suppose that
$X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)\in\textup{M}(V)_U$
is non-cyclic. Exactly one of the following holds:
\begin{itemize}
\item[(i)] $A$ is non-cyclic, or
\item[(ii)] $A$ is cyclic, and $B$ is
non-cyclic, or
\item[(iii)] $A\in\textup{M}(U)$ and $B\in\textup{M}(V/U)$ are cyclic, and $X\in\textup{M}(V)_U$
is non-cyclic.
\end{itemize}
Denote by $n_1,n_2,n_3$ the number of {\it non-cyclic} $X\in\textup{M}(V)_U$
satisfying the pairwise mutually exclusive cases (i), (ii), and (iii),
respectively. The desired probability is $\pi=\pi_1+\pi_2+\pi_3$ where
$\pi_i:=\frac{n_i}{|\textup{M}(V)_U|}$, and $|\textup{M}(V)_U|=q^{n^2-nr+r^2}$.
The cases when $r=1$ or $n-1$ can be handled separately. Suppose $1<r<n-1$.
The probability $\pi_1$ that $A$ is non-cyclic is $\Omega(q^{-3})$
by (\ref{E:NP}). Since the events `$A$ is cyclic' and `$B$~is non-cyclic' are
independent, the probability $\pi_2$ is
$(1-\pi_1)\textup{Prob($B$ non-cyclic)}=\Omega(q^{-3})$.
Explicit upper and lower bounds may be determined by applying (\ref{E:NP}).
The proof is complete once we prove that
$\pi_3=q^{-2}\left(1+\Omega(q^{-1})\right)$.
This is achieved by constructing an upper bound for $n_3$ in
Section~\ref{S:ub}, and a lower bound for $n_3$ in Section~\ref{S:lb}.
\end{proof}
Bounds for the density of non-cyclic matrices in the group $\GL(V)_U$ can be
deduced from those for the density in the algebra $\textup{M}(V)_U$.
Dividing by $|\GL(V)_U|$ instead of $|\textup{M}(V)_U|$ is not problematic
since $|\GL(V)_U|=|\textup{M}(V)_U|\left(1+\Omega(q^{-1})\right)$.
An upper bound for non-cyclic matrices in $\textup{M}(V)_U$
is also an upper bound for non-cyclic matrices in $\GL(V)_U$ as
$\GL(V)_U\subseteq\textup{M}(V)_U$. A lower bound for non-cyclic matrices in $\GL(V)_U$
needs to be altered to ensure that only
{\it invertible} non-cyclic matrices are counted. This requires only
minor modifications to Section~\ref{S:lb}.
Since the \text{\sc Meat-axe}\ is more commonly concerned with algebras and not groups,
we leave this modification to an interested reader.
\medskip\noindent \textbf{Acknowledgements:}\quad The authors are
grateful to Peter Neumann for his advice during many helpful
discussions on this work. The paper grew out of the PhD thesis of
the first author (Brown) undertaken at the University of Western Australia under
the supervision of Giudici and Praeger, and supported by a
University Postgraduate Award. Research for the paper was partially
supported by an Australian Research Council Grant: Giudici and
Praeger are supported by an ARC Australian Research Fellowship
and a Federation Fellowship, respectively.
\section{The upper bound}\label{S:ub}
Let $U$ be an $r$-dimensional subspace of the vector space $V=F^n$ where
$0<r<n$ and $F=\mathbb{F}_q$ is the field with $q$ elements. Let
$\textup{M}(V)_U=\{X\in\textup{M}(V)\mid UX\subseteq U\}$ be the algebra
of $q^{r^2-rn+n^2}$ matrices that normalize $U$. The goal of this section is to
compute an upper bound for the number, $n_3$, of matrices
$X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)\in\textup{M}(V)_U$
for which $U$ is a cyclic $F[A]$-module, $V/U$ is a cyclic $F[B]$-module,
and $V$ is a {\it non-cyclic} $F[X]$-module.
\noindent
\textbf{Notation B.}\quad
As well as Notation A, the following notation will be used in the paper.\par
\hi{$F[t]$ denotes the ring of polynomials with coefficients in $F$;}
\hi{$X\in\textup{M}(V)$ denotes a matrix, and $v\in V$ denotes a (row) vector;}
\hi{$F[X]$ is the subalgebra of $\textup{M}(V)$ comprising all polynomials in $X$
with coefficients in $F$;}
\hi{$vF[X]=\langle v,vX,vX^2,\dots\rangle$ is the {\em cyclic $F[X]$-submodule}
of $V$ generated by $v$;}
\hi{$f\in\Irr(d,F)$ denotes a monic polynomial of degree $d$ which is
irreducible in $F[t]$;}
\hi{$c_X$ denotes the characteristic polynomial $c_X(t)=\det(tI_n-X)$ of
$X\in\textup{M}(n,F)$;}
\hi{$m_X$ denotes the minimal polynomial of $X\in\textup{M}(n,F)$;}
\hi{$V(f)=\{v\in V\mid vf(X)=0\}=\ker f(X)$;}
\hi{$X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)\in\textup{M}(V)_U$
denotes a block matrix with $A\in\textup{M}(r,F)$ and $B\in\textup{M}(n-r,F)$ cyclic,
$C\in F^{(n-r)\times r}$, and $X$ non-cyclic;}
\hi{$\omega(n,q)=\prod_{i=1}^n(1-q^{-i})$; note that $|\GL(n,q)|=q^{n^2}\omega(n,q)$;}
\hi{$C(a)$ the (row) companion matrix of a polynomial
$a(t)=t^r+\sum_{i=0}^{r-1}a_it^i$, see (\ref{E:matform}).}
\setlength{\parskip}{5pt}
There exists a monic irreducible polynomial $f\in\Irr(d,q)$ with
$1\leq d\leq\min(r,n-r)$ for which $V_0:=\ker f(X)$ is a non-cyclic
$F[X]$-module. The restriction, $X_0$, of $X$ to~$V_0$ has
minimal polynomial~$f$. Since $U_0:=V_0\cap U$ and
$(V_0+U)/U\cong V_0/U_0$ are cyclic $F[X]$-modules, it follows that $X_0$
is conjugate to the block diagonal matrix $\textup{diag}(C(f),C(f))$.
The number of $d$-dimensional subspaces $U_0$ of the $r$-dimensional space
$U$ is given by the $q$-binomial coefficient
\begin{equation}\label{E:bin}
\qbinom{r}{d}:=\prod_{i=0}^{d-1}\frac{q^r-q^i}{q^d-q^i}
=q^{d(r-d)}\prod_{i=0}^{d-1}\frac{1-q^{-(r-i)}}{1-q^{-(d-i)}}
=\frac{q^{d(r-d)}\omega(r,q)}{\omega(d,q)\omega(r-d,q)}.
\end{equation}
It is well known that $\qbinom{r}{d}\in\mathbb{Z}[q]$ is a polynomial
in $q$ over~$\mathbb{N}$, and $\deg(\qbinom{r}{d})=d(r-d)$.
First, choose $d$ in the range $1\leq d\leq\min(r,n-r)$, next choose a monic
$f\in\Irr(d,q)$, then a $2d$-dimensional subspace $V_0$ for which
$\dim(V_0\cap U)=\dim((V_0+U)/U)=d$, then choose a linear transformation
$X_0$ on $V_0$ with minimal polynomial $m_{X_0}=f$
satisfying $U_0X_0\subseteq U_0$, and finally choose
an extension $X$ of $X_0$ to $V$. The number of 4-tuples
$(f,V_0,X_0,X)$ overcounts the number~$n_3$ since different $f$
may give the same $X$. Moreover we shall overcount the number of 4-tuples.
In this paragraph the value of $d$ satisfying $1\leq d\leq\min(r,n-r)$ is fixed.
There are at most $\frac{q^d-q}{d}$ choices for $f$ if $d\geq2$, and $q$
choices if $d=1$. How many choices
are there for $V_0$? First, choose $U_0$ in $\qbinom{r}{d}$ ways, then choose
$U+V_0$, or equivalently choose the $d$-dimensional subspace $(U+V_0)/U$
of the $(n-r)$-dimensional space $V/U$ in $\qbinom{n-r}{d}$ ways. Finally,
choose a $d$-dimensional complement $V_0/U_0$ to the $(r-d)$-dimensional
subspace $U/U_0$ in $(U+V_0)/U$ in $q^{d(r-d)}$ ways. Multiplying shows
that there are exactly $\qbinom{r}{d}\qbinom{n-r}{d}q^{d(r-d)}$
choices for $V_0$. Since $U_0X_0\subseteq U_0$ and $X_0$ has minimal
polynomial~$m_{X_0}=f$, it is conjugate in
$\GL(V_0)_{U_0}$ to the $2\times 2$ block diagonal matrix $\textup{diag}(C(f),C(f))$.
The centralizer in $\GL(V_0)_{U_0}$ of $X_0$ has order
$(q^d-1)^2q^d=q^{3d}(1-q^{-d})^2$, and the conjugacy class $X_0^{\GL(V_0)_{U_0}}$
has cardinality
\[
|X_0^{\GL(V_0)_{U_0}}|=\frac{|\GL(V_0)_{U_0}|}{|C_{\GL(V_0)_{U_0}}(X_0)|}=
\frac{q^{3d^2}\omega(d,q)^2}{q^{3d}(1-q^{-d})^2}
=\frac{q^{3(d^2-d)}\omega(d,q)^2}{(1-q^{-d})^2}.
\]
Specifying $X_0$, can be viewed (after a change of basis) as specifying $d$ of
the top $r$ rows, and $d$ of the bottom $n-r$ rows
of $X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)$.
The remaining rows can be completed in at most $|\textup{M}(V)_U|q^{-(r+n)d}$ ways.
This shows
\[
n_3\leq\sum_{d=1}^{\min(r,n-r)} |\Irr(d,q)|\cdot\qbinom{r}{d}\qbinom{n-r}{d}
\frac{q^{d(r-d)}}{1}\cdot
\frac{q^{3(d^2-d)}\omega(d,q)^2}{(1-q^{-d})^2}\cdot
\frac{|\textup{M}(V)_U|q^{-(r+n)d}}{1}.
\]
Equation (\ref{E:bin}) yields $\qbinom{r}{d}\leq\frac{q^{d(r-d)}}{\omega(d,q)}$
and $\qbinom{n-r}{d}\leq\frac{q^{d(n-r-d)}}{\omega(d,q)}$. This, in turn, shows
\[
n_3\leq\sum_{d=1}^{\min(r,n-r)} |\Irr(d,q)|\cdot
\frac{q^{d(r-d)+d(n-r-d)+d(r-d)}}{\omega(d,q)^2}\cdot
\frac{q^{3(d^2-d)}\omega(d,q)^2}{(1-q^{-d})^2}\cdot
\frac{|\textup{M}(V)_U|q^{-(r+n)d}}{1}.
\]
Collecting powers of $q$ gives $q^{-3d}$. Cancelling $\omega(d,q)^2$,
dividing by $|\textup{M}(V)_U|$,
and using the inequality $|\Irr(d,q)|\leq\frac{q^d-q}{d}$ for $d\geq2$ gives
\begin{equation}\label{E:lower}
\begin{aligned}
\frac{n_3}{|\textup{M}(V)_U|}&\leq\frac{q^{-2}}{(1-q^{-1})^2}+
\sum_{d=2}^{\min(r,n-r)} \frac{q^d-q}{d}\cdot
\frac{q^{-3d}}{(1-q^{-d})^2}\\
&<\frac{q^{-2}}{(1-q^{-1})^2}+
\sum_{d=2}^{\infty} \frac{q^{-2d}}{d(1-q^{-d})}.
\end{aligned}
\end{equation}
The first term is $\frac{q^{-2}}{(1-q^{-1})^2}\leq q^{-2}(1+6q^{-1})$, and the
infinite sum is less than $\frac{8q^{-4}}{9}$ because
$\frac{1}{d(1-q^{-d})}\leq\frac{1}{2(1-2^{-2})}\leq\frac23$ for $d\geq2$,
and $\sum_{d=2}^\infty q^{-2d}=\frac{q^{-4}}{1-q^{-2}}\leq\frac{4q^{-4}}{3}$. Hence
\[
\frac{n_3}{|\textup{M}(V)_U|}<q^{-2}\left(1+6q^{-1}\right)
+\frac{4q^{-3}}{9}\leq q^{-2}\left(1+\frac{58q^{-1}}{9}\right).
\]
\section{Counting polynomials}\label{S:Polys}
The goal of this section is to prove a simple combinatorial
result for polynomials over~$\mathbb{F}_q$. This result will be used in
Section~\ref{S:lb} to prove a lower bound for $n_3$.
Morrison~\cite{M} proves that the density of coprime pairs of polynomials
of degree {\em at most}~$r$ over $\mathbb{F}_q$ is $1-q^{-1}+q^{-2r-1}-q^{-2r-2}$.
A simpler answer exists if the degrees are {\em precisely}~$r$.
\begin{lemma}\label{L:coprime}
Let $\mathcal{M}_r$ denote the set of $q^r$ monic polynomials in $\mathbb{F}_q[t]$ of
degree~$r$.
\begin{itemize}
\item[(a)] The number of coprime ordered pairs $(a,b)$ in $\mathcal{M}_r\times\mathcal{M}_s$
is $q^{r+s}(1-q^{-1})$ when $rs>0$, and $q^{r+s}$ when $rs=0$.
\item[(b)] Fix $f\in\Irr(d,q)$ and suppose $1\leq d\leq\min(r,s)$. Then the
number of coprime pairs $(a,b)$ in $\mathcal{M}_r\times\mathcal{M}_s$ satisfying
$\gcd(f,ab)=1$ is at least $q^{r+s}(1-q^{-1}-2q^{-d}+2q^{-2d})$.
\end{itemize}
\end{lemma}
\begin{proof}
(a) Let $c(r,s)$ denote the number of coprime ordered pairs
$(a,b)\in\mathcal{M}_r\times\mathcal{M}_s$. The cardinality of $\mathcal{M}_r\times\mathcal{M}_s$,
{\em viz.} $|\mathcal{M}_r|\,|\mathcal{M}_s|=q^{r+s}$, can be determined
in a different way.
An ordered pair $(a,b)\in\mathcal{M}_r\times\mathcal{M}_s$ has $\gcd(a,b)=d$
if and only if $\gcd(\frac ad,\frac bd)=1$. If $\deg(d)=k$, then there
are $q^k$ choices for $d$, and $c(r-k,s-k)$ pairs $(\frac ad,\frac bd)$. Thus
\[
|\mathcal{M}_r\times\mathcal{M}_s|=\sum_{k=0}^{\min(r,s)} |\mathcal{M}_k|\,c(r-k,s-k),
\quad\textup{or}\quad
q^{r+s}=\sum_{k=0}^{\min(r,s)} q^k c(r-k,s-k).
\]
Rearranging gives a recurrence relation
$c(r,s)=q^{r+s}-\sum_{k=1}^{\min(r,s)} q^k c(r-k,s-k)$
with initial conditions $c(r,0)=q^r$, $c(0,s)=q^s$. Induction may be used to
prove $c(r,s)=q^{r+s}(1-q^{-1})$ holds when $rs>0$. (The sum in the recurrence
telescopes to $q^{r+s-1}$.) It is noteworthy that the
probability $c(r,s)/q^{r+s}=1-q^{-1}$ is independent of both $r$ and $s$.
(b) Assume $f\in\Irr(d,q)$ and $1\leq d\leq\min(r,s)$.
We shall underestimate the number of coprime ordered pairs
$(a,b)\in\mathcal{M}_r\times\mathcal{M}_s$ for which $\gcd(ab,f)=1$. By part~(a) there
are $q^{r+s}(1-q^{-1})$ coprime pairs $(a,b)\in\mathcal{M}_r\times\mathcal{M}_s$.
The number of $a\in\mathcal{M}_r$ divisible by $f$ is $q^{r-d}$, and the number of
$(a,b)\in\mathcal{M}_r\times\mathcal{M}_s$ with $f\mid a$ and $f\mid b$ is $q^{r+s-2d}$.
Hence $q^{r+s-d}-q^{r+s-2d}$ ordered pairs $(a,b)$ have $f\mid a$ and
$f\nmid b$. The same count holds for ordered pairs $(a,b)$ with
$f\nmid a$ and $f\mid b$. However, some of these ordered pairs may
not be coprime, and therefore
$q^{r+s}(1-q^{-1})-2(q^{r+s-d}-q^{r+s-2d})$ underestimates the number of
coprime $(a,b)$ with $\gcd(f,ab)=1$. Rearranging proves the result.
\end{proof}
A heuristic argument suggests that the matrices
$X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)\in\textup{M}(V)_U$
for which $c_A$ and $c_B$ are not coprime, has density roughly
$q^{-1}$. An extra factor of $q^{-1}$ arises when we insist that
$X$ is non-cyclic. This is basically because there are $q$
non-cyclic matrices in $\textup{M}(V)_U$ when $\dim(V)=2$ and $\dim(U)=1$,
as $C$ must be 0. A rigorous argument is given below.
\section{The lower bound}\label{S:lb}
Fix $X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)\in\textup{M}(V)_U$.
Then $V$ becomes an $F[t]$-module with $v*f(t)=vf(X)$ where the juxtaposition
$vf(X)$ denotes vector-times-matrix multiplication.
We also say that~$V$ is an $F[X]$-module, where $F[X]$ is the subalgebra
of $\textup{M}(V)_U$ comprising all polynomials in $X$ over $F$.
In this section we give a lower bound for $n_3$ by underestimating
the number of matrices $X\in\textup{M}(V)_U$
which have a {\em unique} non-cyclic primary submodule. For these matrices,
$U$ is a cyclic $F[A]$-module, $V/U$ is a cyclic $F[B]$-module, and
$V$ is a {\it non-cyclic} $F[X]$-module. That is, we are counting certain
$X=\left(\begin{smallmatrix}A&0\\textup{C}&B\end{smallmatrix}\right)\in\textup{M}(V)_U$
for which $c_A=m_A$, $c_B=m_B$, and $c_X=c_Ac_B\neq m_X$.
Since $U$ and $V/U$ are cyclic, there exist vectors
$u\in U$ and $v+U\in V/U$, generating the respective $F[t]$-modules.
Consider the basis
\begin{equation}\label{E:X}
u,uX,\dots,uX^{r-1},v,vX,\dots,vX^{n-r-1}
\end{equation}
for $V$. Then $X$ is conjugate in $\GL(V)_U$ to a matrix of the form
$\left(\begin{smallmatrix}A'&0\\textup{C}'&B'\end{smallmatrix}\right)\in\textup{M}(V)_U$
where
\begin{equation}\label{E:matform}
A'=\left(\begin{smallmatrix}
0&1& &0\\&&\ddots&\\0&0& &1\\-a_0&-a_1&\cdots&-a_{r-1}
\end{smallmatrix}\right),\quad
B'=\left(\begin{smallmatrix}
0&1& &0\\&&\ddots&\\0&0& &1\\-b_0&-b_1&\cdots&-b_{n-r-1}
\end{smallmatrix}\right),\quad
C'=\left(\begin{smallmatrix}0&0&\cdots&0\\\vdots&\vdots&&\vdots\\0&0&\cdots&0\\
c_0&c_1&\cdots&c_{r-1}\end{smallmatrix}\right).
\end{equation}
Set $a:=t^r+\sum_{i=0}^{r-1}a_it^i=m_A$,
$b:=t^{n-r}+\sum_{i=0}^{n-r-1}b_it^i=m_B$, and
$c:=\sum_{i=0}^{r-1}c_it^i$.
Then $ua(X)=0$ and $vb(X)=uc(X)$ where $\deg(c)<\deg(a)$.
The matrices $A'$ and $B'$ are called {\it companion matrices} of~$a$
and~$b$ and are abbreviated $C(a)$ and $C(b)$, respectively.
We shall count non-cyclic matrices $X$ for which $a=fg$, $b=fh$,
$f\in\Irr(d,q)$, and $\gcd(f,gh)=\gcd(g,h)=1$.
Note that $V=V(f)\oplus V(gh)$ where $X$ is non-cyclic on $V(f):=\ker f(X)$,
and cyclic on $V(gh)=V(g)\oplus V(h)$. Such matrices $X$ are conjugate in
$\GL(V)_U$ to the block diagonal matrix $X_{f,g,h}:=\textup{diag}(C(g),C(f),C(f),C(h))$
for a uniquely determined triple $(f,g,h)$. This fact is needed to establish
a lower
bound for~$n_3$. (Different choices for~$f$ give
different $X_{f,g,h}$ due to our assumption that $V(f)$ is the {\em unique}
non-cyclic primary $F[X]$-submodule of $V$.)
As $X$ is conjugate
in $\GL(V)$ to $\textup{diag}(C(f)\oplus C(f),C(gh))$, it follows that
$|C_{\GL(V)_U}(X)|\leq q^{3d}(1-q^{-d})^2(q^{n-2d}-1)$ because
\[
C_{\GL(V)_U}(X)\leq C_{\GL(V)}(X)
\cong C_{\GL(V(f))}(C(f)\oplus C(f))\times C_{\GL(V(gh))}(C(gh)).
\]
First, choose $d$ in the range $1\leq d\leq\min(r,n-r)$, next choose a monic
$f\in\Irr(d,q)$, then choose an ordered pair $(g,h)$ satisfying
$\gcd(f,gh)=\gcd(g,h)=1$. By Lemma~\ref{L:coprime}(b), there are at least
\[
q^{(r-d)+(n-r-d)}(1-q^{-1}-2q^{-d}+2q^{-2d})=q^{n-2d}(1-q^{-1}-2q^{-d}+2q^{-2d})
\]
ordered pairs $(g,h)$. Summing over the relevant triples $(f,g,h)$ gives
\[
n_3\geq\sum_{d=1}^{\min(r,n-r)}\kern-8pt\sum_{f\in\Irr(d,q)}\;\sum_{(g,h)}
\frac{|\GL(V)_U|}{|C_{\GL(V)_U}(X_{f,g,h})|}.
\]
But $|\GL(V)_U|=|\textup{M}(V)_U|\omega(r,q)\omega(n-r,q)$ and
$|C_{\GL(V)_U}(X)|\leq q^{3d}(1-q^{-d})^2(q^{n-2d}-1)$~so
\[
\frac{n_3}{|\textup{M}(V)_U|}\geq\sum_{d=1}^{\min(r,n-r)}
|\Irr(d,q)|\frac{\omega(r,q)\omega(n-r,q)}{q^{3d}(1-q^{-d})^2(q^{n-2d}-1)}
\cdot q^{n-2d}(1-q^{-1}-2q^{-d}+2q^{-2d}).
\]
Euler's pentagonal number theorem shows that
$\omega(\infty,q)>1-q^{-1}-q^{-2}+q^{-5}$.
Therefore
\begin{equation}\label{E:LB}
\frac{n_3}{|\textup{M}(V)_U|}\geq\sum_{d=1}^{\min(r,n-r)} |\Irr(d,q)|
\frac{q^{-3d}(1-q^{-1}-q^{-2}+q^{-5})^2}{(1-q^{-d})^2(1-q^{-(n-2d)})}
\cdot (1-q^{-1}-2q^{-d}+2q^{-2d}).
\end{equation}
The number $n_3$ depends on $r$. To emphasize this dependence we
write $n_3(r)$. The automorphism of $\textup{M}(V)$ obtained by conjugating by
$e_i\leftrightarrow e_{n-i}$ and then transposing, swaps the maximal reducible
algebras $\textup{M}(V)_{U(r)}$ and $\textup{M}(V)_{U(n-r)}$. Hence $n_3(r)=n_3(n-r)$.
By swapping $r$ and $n-r$, if necessary, we shall assume that
$\min(r,n-r)=r$. It is convenient to give a sharper lower bound
than (\ref{E:LB}) in the case that $r=1$. The calculation above
has $a=t-\lambda=f$, $g=1$, and $b=fh$ where $h(\lambda)\neq0$.
There are precisely $q^{n-2}(1-q^{-1})$ choices for $h$. (This is
a sharper estimate than given above.) Hence when $r=1$, we have $d=1$.
Since $|\Irr(1,q)|=q$ and
$1-q^{-1}-q^{-2}+q^{-5}=(1-q^{-1})(1-q^{-2}-q^{-3}-q^{-4})$,
a sharper bound than (\ref{E:LB}) for $n\geq3$ is
\begin{equation}\label{E:pi3lb}
\begin{aligned}
\frac{n_3(1)}{|\textup{M}(V)_U|}&\geq
\frac{q\cdot q^{-3}(1-q^{-1}-q^{-2}+q^{-5})^2q^{n-2}(1-q^{-1})}
{(1-q^{-1})^2(q^{n-2}-1)}\\
&\geq q^{-2}(1-q^{-2}-q^{-3}-q^{-4})^2(1-q^{-1}).
\end{aligned}
\end{equation}
This bound also holds when $n=2$ and $r=1$, as a direct calculation shows
that $\frac{n_3(1)}{|\textup{M}(V)_U|}=q^{-2}$ in this case.
Henceforth assume that $r\geq2$, and hence that $n\geq4$.
The summand in (\ref{E:LB}) with $d=1$ is greater than
\begin{equation}\label{E:d=1}
q^{-2}(1-q^{-2}-q^{-3}-q^{-4})^2(1-3q^{-1}+2q^{-2})\geq q^{-2}(1-3q^{-1}+4q^{-3}).
\end{equation}
It follows from (\ref{E:LB}) and (\ref{E:d=1}) that
\begin{equation}\label{E:lb}
\frac{n_3(r)}{|\textup{M}(V)_U|}\geq q^{-2}(1-3q^{-1}+4q^{-3})
\geq q^{-2}\left(1-2q^{-1}\right)
\end{equation}
holds for $r\geq2$. However, the bound (\ref{E:lb}) when $r\geq2$ is always
smaller than the bound~(\ref{E:pi3lb}) when $r=1$. Thus (\ref{E:lb}) gives
a uniform lower bound for all $r$ satisfying $0<r<n$.
\begin{proof}[Proof of Theorem~\textup{\ref{T:Main}}]
Recall the notation $n_i$ and $\pi_i=\frac{n_i}{|\textup{M}(V)_U|}$ used in
the `Proof Strategy' in Section~\ref{S:Main}. We shall prove
$q^{-2}(1+c_1q^{-1})\leq\pi\leq q^{-2}(1+c_2q^{-1})$, where
$\pi=\pi_1+\pi_2+\pi_3$ equals $\textup{Prob($X\in\textup{M}(V)_U$ is non-cyclic)}$,
and $c_1=-\frac{4}{3}$ and $c_2=\frac{35}{3}$.
As mentioned previously,
we shall assume that $\min(r,n-r)=r$. If $n=2$, then only the $q$ scalar
matrices of the $q^3$ elements of $\textup{M}(V)_U$ are non-cyclic. Thus we have
$\pi=\pi_3=q^{-2}$ and the stated bounds
$q^{-2}(1-\frac{4q^{-1}}{3})\leq q^{-2}\leq q^{-2}(1+\frac{35q^{-1}}{3})$
clearly hold. Suppose now that $n\geq3$. Consider
the case when $r=1$. Then the probability $\pi_1$ that $A$ is non-cyclic is $0$,
and $\frac{2q^{-3}}{3}\leq\pi_2\leq\frac{8q^{-3}}{3}$ by (\ref{E:NP}) because
$n-r\geq2$. We have shown in Section~\ref{S:ub}, and above, that
\begin{equation}\label{E:pi3}
q^{-2}\left(1-2q^{-1}\right)\leq\pi_3
\leq q^{-2}\left(1+\frac{58q^{-1}}{9}\right)
\qquad\textup{for $n\geq1$.}
\end{equation}
Adding $\pi_1=0$ and $\frac{2q^{-3}}{3}\leq\pi_2\leq\frac{8q^{-3}}{3}$ and
$q^{-2}-2q^{-3}\leq\pi_3 \leq q^{-2}+\frac{58q^{-3}}{9}$ gives
$q^{-2}\left(1-\frac{4q^{-1}}{3}\right)\leq\pi\leq q^{-2}\left(1+\frac{82q^{-1}}{9}\right)$
when $r=1$.
Now consider the case when $2\leq r\leq n-r$. Then
$\frac1{12}\leq\frac{2q^{-3}}{3}\leq\pi_1\leq\frac{8q^{-3}}{3}\leq\frac13$
holds by~(\ref{E:NP}). However, $\pi_2$ equals $1-\pi_1$ times the probability
that $B$ is non-cyclic, and hence
\[
\frac{4q^{-3}}{9}\leq\left(1-\pi_1\right)\frac{2q^{-3}}{3}\leq
\pi_2\leq\left(1-\pi_1\right)\frac{8q^{-3}}{3}
\leq\frac{88q^{-3}}{36}.
\]
Adding $\frac{2q^{-3}}{3}\leq\pi_1\leq\frac{8q^{-3}}{3}$ and
$\frac{4q^{-3}}{9}\leq\pi_2\leq\frac{88q^{-3}}{36}$ to the bounds (\ref{E:pi3})
for $\pi_3$ gives
\[
q^{-2}\left(1-\frac{8q^{-1}}{9}\right)
\leq\pi_1+\pi_2+\pi_3\leq q^{-2}\left(1+\frac{104q^{-1}}{9}\right)
\qquad\textup{for $r\geq2$.}
\]
The constants $c_1=-\frac{4}{3}$ and $c_2=\frac{35}{3}$ suffice
as $-\frac{4}{3}<-\frac{8}{9}$ and $\frac{82}{9}<\frac{104}{9}<\frac{35}{3}$.
\end{proof}
|
\section{Introduction}
\label{sec:intro}
The experimental community is engaged in an effort to characterize the properties of
dark matter using nuclear recoil as the signal for the elastic scattering of heavy dark
matter particles off nuclear targets. While astrophysics has established certain basic properties of
dark matter -- that it is long-lived or stable, cold or warm, gravitationally active, but without
strong couplings to itself or to baryons -- the nature of the interactions of
dark-matter particles with ordinary matter is otherwise unknown \cite{Freese2013,Bertone2005}. Among the leading
dark-matter candidates are weakly interacting massive particles (WIMPs): it was
noted long ago that such weak-scale particles could be produced in the Big Bang
with abundances on the order of that required by observation, $\Omega_{DM} \sim 0.27$.
Additional motivation for WIMPs comes from expectations that new physics at the
mass-generation scale of the standard model might naturally lead to
new particles of mass $\sim$ 10 GeV - 10 TeV.
One of the promising methods for detecting WIMPs is via their elastic
scattering with heavy nuclei (``direct detection") \cite{Goodman1985,Drukier1986,Jungman1996}.
The WIMPs would be slowly moving today, with velocities $v_T \sim 10^{-3}$,
and consequently with momenta $q \sim 100$ MeV/c and kinetic energies
$\epsilon \sim$ 50 keV.
The momentum scale is on the order of the inverse size of the nucleus,
$\hbar /R_\mathrm{nucleus} \sim \hbar /(1.2 fm~ A^{1/3}) \sim 30$ MeV/c, while $\epsilon$
is small compared to the typical excitation energies in a nucleus of $\sim$ 1 MeV.
Consequently
\begin{itemize}
\item the detection of WIMP elastic interactions with a nucleus
requires sensitivity to recoil energies on the order of 10s of keV;
\item inelastic interactions need only be considered in unusual cases where a target
nucleus has an excited state within $\sim$ 100 keV of the ground state; and
\item a proper quantum mechanical treatment of the elastic scattering cross section
should take into account the size of the nucleus, as $q R_\mathrm{nucleus} \gtrsim 1$.
\end{itemize}
Because the WIMP will, in most cases, only scatter elastically, one also sees that
parity and time-reversal selection rules that operate for diagonal matrix elements
will limit what can be learned in direct detection experiments.
While we know little about dark matter interactions with ordinary matter,
their possible associations with electroweak interactions suggests using the standard model as a guide.
In electromagnetism, elastic scattering can
occur through charge or magnetic interactions. Both interactions involve nontrivial isospin --
the charge coupling is only to protons, while the magnetic coupling involves the distinct proton
and neutron magnetic moments. Magnetic elastic scattering occurs through two interfering
three-vector operators, spin $\vec{\sigma}(i)$ and orbital angular momentum $\vec{\ell}(i)$.
For weak interactions, the weak charge operator couples primarily to neutrons, while the axial-charge
operator $\vec{\sigma}(i) \cdot \vec{p}(i)$ makes effectively no contribution to elastic scattering, apart from small recoil corrections,
due to the constraints imposed by parity and time-reversal invariance. One might expect,
consequently, that the WIMP-nuclear interaction will involve a variety of operators
as well as couplings that depend on isospin.
In part for historical reasons, WIMP elastic scattering experiments are most often
analyzed by assuming the interaction is simpler than those described above: isoscalar,
coupled either to the nucleon number operator $1(i)$ (spin-independent or SI) or the
nucleon spin $\sigma(i)$ (spin-dependent or SD) \cite{Goodman1985,Smith1990,Lewin1996}.
These are the operators for a point nucleus. While a form factor is often introduced to account
phenomenologically for the fact that the momentum transfer is large on the nuclear scale,
the quantum mechanical consequences of $o(1)$ operators like $\vec{q} \cdot \vec{r}(i)$
have been largely neglected.
Recently there have been efforts to treat the WIMP-nucleon interaction in more generality,
using the tools of effective field theory (EFT) \cite{Fan2010,Liam2013,Hill2013,Nikhil2014}. We
describe the approach of \cite{Liam2013,Nikhil2014} in Sec. \ref{sec:response} and
its consequences for WIMP-nucleus elastic scattering. Consistent with general symmetry
arguments, six independent nuclear response functions are identified, in contrast to the
two assumed in SI/SD treatments. The new operators are associated with derivative couplings,
where a proper treatment of $\vec{q} \cdot \vec{r}(i)$ is essential due to the need to identify
associated parity- and time-reversal-conserving elastic operators. When this is done, we find
that velocity-dependent interactions lead to cross sections
$\sim q^2/m_N^2 ~G_F^2\sim 10^{-2} ~G_F^2$, where $m_N$
in the nucleon mass and $G_F$ the weak coupling constant, in contrast to the SI/SD result,
$\sim v_T^2~ G_F^2 \sim 10^{-6} ~G_F^2$. Our effective theory treatment
shows that much more can be learned about the properties of WIMP dark matter from
elastic scattering experiments than is generally appreciated. However, it also shows that a
greater variety of experiments will be necessary to extract this information and to
eliminate possible sources of confusion, when competing experiments are compared.\\
\section{The Nuclear Elastic Response from Effective Theory}
\label{sec:response}
Here we summarize the effective theory construction of the WIMP-nucleon interaction of Ref. \cite{Liam2013,Nikhil2014}.
Details can be found in the original papers.
The Lagrangian density for the scattering of a WIMP off a nucleon is taken to have the form
\begin{equation}
\mathcal{L}_{\textrm{int}}(\vec{x}) =c ~ \Psi^*_\chi(\vec{x}) {{\mathcal{O}}}_\chi \Psi_\chi(\vec{x})~ \Psi^*_N(\vec{x}) {{\mathcal{O}}}_N \Psi_N(\vec{x}),
\end{equation}
where the $\Psi(\vec{x})$ are nonrelativistic fields and where the WIMP and
nucleon operators ${{\mathcal{O}}}_\chi$ and $ {{\mathcal{O}}}_N$ may have vector indices.
The operators ${{\mathcal{O}}}_\chi$ and ${{\mathcal{O}}}_N$ are then allowed to take on their most
general form, constrained by imposing relevant symmetries.
The construction was done in the nonrelativistic limit to second order in the momenta.
Thus the relevant operators are those appropriate for use with
Pauli spinors. The Galilean-invariant amplitudes take the form
\begin{equation}
\sum_{i=1}^{\cal{N}} \left( c_i^\mathrm{n} ~{{\mathcal{O}}}_i^{\mathrm{\,n}} + c_i^{\mathrm{p}}~{{\mathcal{O}}}_i^{\mathrm{\,p}} \right),
\label{eq:Lag}
\end{equation}
where the coupling coefficients $c_i$ may be different for proton and neutrons.
The number $\cal{N}$ of such operators $\mathcal{O}_i$ -- which have the product form $\mathcal{O}^i_\chi \otimes\mathcal{O}^i_N$ -- depends
on the generality of the particle physics description.
The WIMP and nucleon operators for this construction include $1_\chi$ and
$1_N$, the three-vectors $\vec{S}_\chi$ and $\vec{S}_N$, and the momenta of the WIMP and nucleon. Of the four momenta involved in the scattering (two incoming and two outgoing), only two combinations are physically relevant due to inertial frame-independence and momentum conservation. It is convenient to work with the frame-invariant quantities, the momentum transfer $\vec{q}$ and the WIMP-nucleon
relative velocity,
\begin{equation}
\vec{v} \equiv \vec{v}_{\chi, \rm in} - \vec{v}_{N, \rm in}.
\end{equation}
The Hermitian form of this velocity is
\begin{equation}
\vec{v}^\perp = \vec{v} + \frac{\vec{q}}{2\mu_N} = {1 \over 2} \left( \vec{v}_{\chi,\mathrm{in}}+\vec{v}_{\chi,\mathrm{out}} - \vec{v}_{N,\mathrm{in}} - \vec{v}_{N,\mathrm{out}} \right) =
{1 \over 2} \left( {\vec{p} \over m_\chi}+{\vec{p}^{\, \prime} \over m_\chi} - {\vec{k} \over m_N} - {\vec{k}^{\, \prime} \over m_N} \right)
\label{eq:vperp}
\end{equation}
which satisfies $\vec{v}^\perp \cdot \vec{q} =0$ as a consequence of energy conservation. Here $\mu_N$ is the WIMP-nucleon reduced mass, $\vec{p}$ ($\vec{p}^{\,\prime}$)
is the incoming (outgoing) WIMP three-momentum, and $\vec{k}$ ($\vec{k}^\prime$) the incoming
(outgoing) nucleon three-momentum.
It was shown in \cite{Liam2013} that operators are guaranteed to be Hermitian if they are built out of the following four three-vectors,
\begin{equation}
\vec{S}_\chi~~~~~~~ \vec{S}_N~~~~~~~\vec{v}^\perp~~~~~~~i {\vec{q} \over m_N}
\end{equation}
The nucleon mass $m_N$ has been introduced as a convenient scale to render $\vec{q}/m_N$ and the constructed
${{\mathcal{O}}}_i$ dimensionless: the choice of this scale is not arbitrary, but is instead connected with the velocity
operator when that operator is embedded in a nucleus, as we discuss later. One finds to second order in velocities and
momenta,
\begin{eqnarray}
{\mathcal{O}}_1 &=& 1_\chi 1_N \nonumber \\
{\mathcal{O}}_2 &=& (v^\perp)^2 \nonumber \\
{\mathcal{O}}_3 &=& i \vec{S}_N \cdot ({\vec{q} \over m_N} \times \vec{v}^\perp) \nonumber \\
{\mathcal{O}}_4 &=& \vec{S}_\chi \cdot \vec{S}_N \nonumber \\
{\mathcal{O}}_5 &=& i \vec{S}_\chi \cdot ({\vec{q} \over m_N} \times \vec{v}^\perp) \nonumber \\
{\mathcal{O}}_6&=& (\vec{S}_\chi \cdot {\vec{q} \over m_N}) (\vec{S}_N \cdot {\vec{q} \over m_N}) \nonumber \\
{\mathcal{O}}_7 &=& \vec{S}_N \cdot \vec{v}^\perp \nonumber \\
{\mathcal{O}}_8 &=& \vec{S}_\chi \cdot \vec{v}^\perp \nonumber \\
{\mathcal{O}}_9 &=& i \vec{S}_\chi \cdot (\vec{S}_N \times {\vec{q} \over m_N}) \nonumber \\
{\mathcal{O}}_{10} &=& i \vec{S}_N \cdot {\vec{q} \over m_N} \nonumber \\
{\mathcal{O}}_{11} &=& i \vec{S}_\chi \cdot {\vec{q} \over m_N} \nonumber \\
{\mathcal{O}}_{12} &=& \vec{S}_\chi \cdot (\vec{S}_N \times \vec{v}^\perp) \nonumber \\
{\mathcal{O}}_{13} &=&i (\vec{S}_\chi \cdot \vec{v}^\perp ) ( \vec{S}_N \cdot {\vec{q} \over m_N}) \nonumber \\
{\mathcal{O}}_{14} &=& i ( \vec{S}_\chi \cdot {\vec{q} \over m_N})( \vec{S}_N \cdot \vec{v}^\perp ) \nonumber \\
{\mathcal{O}}_{15} &=& - ( \vec{S}_\chi \cdot {\vec{q} \over m_N}) ((\vec{S}_N \times \vec{v}^\perp) \cdot {\vec{q} \over m_N})
\label{eq:ops}
\end{eqnarray}
The first eleven of these operators can be generated by interactions
involving only spin-0 or spin-1 mediators, and were
discussed in \cite{Liam2013}. One of these,
${\mathcal{O}}_2$, was later eliminated from consideration in \cite{Nikhil2014}, as it cannot be
obtained from the leading-order non-relativistic reduction
of a manifestly relativistic operator. The remaining operators are associated with more
complicated exchanges. These operators can be labeled
as LO, NLO, and N$^2$LO, according to the total number of
momenta and velocities they contain,
corresponding to total cross sections that scale as $v_T^0$, $v_T^2$, or $v_T^4$, where $v_T$ is the WIMP
velocity in the laboratory frame (though, below, we stress that this scaling is accompanied by
additional factors that arise from the treatment of velocity-dependent interactions). Operator ${\mathcal{O}}_{15}$ is
cubic in velocities and momenta, generating a total cross section of order $v_T^6$ (N$^3$LO).
It is retained because it arises as the leading-order nonrelativistic limit of a convariant
four-fermion interaction, as discussed in \cite{Nikhil2014}.
Each operator
can have distinct couplings to protons and neutrons. Introducing isospin, which
is also useful as an approximate symmetry of the nuclear wave functions, the general form
of the effective theory interaction becomes
\begin{equation}
\sum_{i=1}^{15} ( c_i^{0} 1 + c_i^{1} \tau_3) {\mathcal{O}}_i = \sum_{\tau=0,1} \sum_{i=1}^{15} c_i^\tau {\mathcal{O}}_i t^\tau,
~~~ c_2^\tau \equiv 0, \nonumber \\
\label{eq:Lagrangian}
\end{equation}
where $c_i^{0} = {1 \over 2} (c_i^{\mathrm{p}}+c_i^{\mathrm{n}})$, $c_i^{1} = {1 \over 2} (c_i^{\mathrm{p}}-c_i^{\mathrm{n}})$, $t^0 = 1$, and
$t^1=\tau_3$.
Thus the EFT has a total of 28 parameters, associated with 14 space/spin operators each
of which can have distinct couplings to protons and neutrons.
This interaction can then be embedded in the nucleus. The procedure followed in
\cite{Liam2013,Nikhil2014} assumes that the nuclear interaction is the
sum of the WIMP interactions with the individual nucleons in the nucleus. The nuclear
operators then involve a convolution of the ${\mathcal{O}}_i$, whose momenta must now be treated
as local operators appropriate for bound nucleons, with the plane wave associated with
the WIMP scattering, which is an angular and radial operator that can be decomposed
with standard spherical harmonic methods. Because momentum transfers are typically
comparable to the inverse nuclear size, it is crucial to carry through such a multipole
decomposition in order to identify the nuclear responses associated with the various
$c_i$s. The scattering probability is given by the square of the (Galilean) invariant
amplitude $\mathcal{M}$, a product of WIMP and nuclear matrix elements, averaged over initial
WIMP and nuclear magnetic quantum numbers $M_\chi$ and $M_N$, and summed over final
magnetic quantum numbers.
The result can be organized
in a way that factorizes the particle and nuclear physics
\begin{equation}
\frac{1}{2j_\chi + 1}\frac{1}{2j_N+1}\sum_{\textrm{spins}} |\mathcal{M}|^2 \equiv \sum_{k}
\sum_{\tau,\tau^\prime=0}^1 R_k \left( {v}_T, {{q}^{\,2} \over m_N^2},\left\{c_i^\tau c_j^{\tau^\prime} \right\} \right)
~W_k^{\tau \tau^\prime}( {q}^{\,2} b^2)
\label{eq:RS}
\end{equation}
where the sum extends over products of WIMP response functions $R_k$ and
nuclear response functions $W_k$. The $R_k$ isolate the particle physics: they depend on
specific combinations of bilinears in the low-energy constants (LECs) of the EFT -- the 2$\cal{N}$
coefficients of Eq. (\ref{eq:Lag}) -- labeled by isospin $\tau$ (isoscalar, isovector). The WIMP response functions also depend
on the magnitude of the WIMP velocity ${v}_T$, measured relative to the center of mass of the
nucleus, and the three-momentum transfer
$\vec{q}= \vec{p}^{\,\prime}-\vec{p}=\vec{k} - \vec{k}^{\prime}$.
The nuclear response functions $W_k$ can be varied by experimentalists, if they explore a variety
of nuclear targets.
The $W_k$ are functions of $y \equiv (q b/2)^2$, where $b$ is the nuclear size (explicitly the
harmonic oscillator parameter if the nuclear wave functions are expanded in that
single-particle basis).
The embedding of the WIMP-nucleon interaction in the nucleus will necessarily lead to the most
general form of the WIMP-nucleus interaction, provided the quantum mechanics is done properly.
But it is also satisfying to derive the end result from symmetry considerations alone. Our EFT
operators are constructed from the nucleon charge, spin, and velocity. These operators can
be combined to form a nucleon vector charge operator $1(i)$, an axial-vector charge $\vec{\sigma}(i) \cdot \vec{v}(i)$,
an axial spin current $\vec{\sigma}(i)$, a vector velocity current $\vec{v}(i)$, and a vector
spin-velocity current $\vec{\sigma}(i) \times \vec{v}(i)$. Charges and currents can be combined with
the plane-wave operator appearing in the scattering,
$ e^{i \vec{q} \cdot \vec{r}(i)} $,
to produce nuclear multipoles operators $\hat{O}_J$ carrying definite angular momentum and parity, and
transforming with a definite sign under time reversal. In the case of the currents, one can take
three projections -- longitudinal, transverse electric, and transverse magnetic -- relative to a
quantization axis defined by the momentum transfer.
As the nuclear ground state is nearly an eigenstate of both parity and time reversal, only those multipole
operators that are even under parity and time reversal can contribute. By eliminating all operators with
the wrong transformation properties (a exercise familiar from standard weak interactions theory) one
obtains the surviving response functions and thus the general form of the cross section. The
results, shown in Fig. \ref{fig:fig1}, indicate six response functions are needed to describe WIMP-nucleus
elastic scattering, not the two (SI/SD) usually assumed.
\begin{figure}[t]
\includegraphics[width=160mm]{tabfig.pdf}
\caption{Constraints imposed by the parity and time-reversal properties of the nuclear ground state on the multipole operators
mediating WIMP-nucleus elastic scattering. Transitions forbidden by parity (time reversal) are tinted red (blue); those forbidden
by both are indicated by the red-blue gradient fill.
Six types of multipole operators are allowed, corresponding to the even-angular-momentum-J multipoles of the vector charge operator (which generate the SI response),
the odd-J longitudinal and transverse electric multipoles of the axial spin operator (which generate in combination
the standard SD response), the odd-J transverse
magnetic multipoles of the vector velocity operator, and the even-J longitudinal and transverse electric multipoles of
the vector spin-velocity operator.}
\label{fig:fig1}
\end{figure}
Specific forms for the WIMP and nuclear response functions appearing in Eq. (\ref{eq:RS}) can be obtained from the
EFT Lagrangian, as detailed
in Ref. \cite{Nikhil2014}, under the assumption that the WIMP-nucleus interaction is the sum over the various EFT WIMP-nucleon
interactions. As the arguments just given above indicate there must be six types of nuclear multipole operators
in a model-independent formulation, and as our effective theory Lagrangian gives the most general WIMP-nucleon
interaction, one might anticipate that a straight-forward application of that Lagrangian would then generate the explicit forms
of the nuclear multipole operators. While this takes some algebra, indeed this is what happens. The nuclear
response functions are generated by six operators that we can write, for simplicity, in their long-wavelength
forms (that is, without accompanying form factors). One finds the leading-order multipoles
\begin{eqnarray}
C_{J=0M} \rightarrow \sum_{i=1}^A 1(i)~~~~~~~~~~~~~~~L_{J=1M}^5 \sim T_{J=1M}^{\mathrm{el}~5} \rightarrow \sum_{i=1}^A \sigma_{1M}(i)~~~~~~~~~~~~~~~T_{J=1M}^\mathrm{mag} \rightarrow {q \over m_N} \sum_{i=1}^A \ell_{1 M}(i) ~~~~~~~~~~~~~~~~~~~~\nonumber \\
L_{J=0M} \rightarrow {q \over m_N} \sum_{i=1}^A \vec{\sigma}(i) \cdot \vec{\ell}(i) ~~~~~~~~ L_{J=2M} \rightarrow {q \over m_N} \sum_{i=1}^A \left[ r(i) \otimes \left( \vec{\sigma}(i) \times {1 \over i} \vec{\nabla} \right) \right]_{2M} ~~~~~~~~
T^\mathrm{el}_{J=2M} \rightarrow {q \over m_N} \sum_{i=1}^A \left[ r(i) \otimes \left( \vec{\sigma}(i) \times {1 \over i} \vec{\nabla} \right) \right]_{2M} \nonumber
\end{eqnarray}
The first two operators are the standard SI/SD interactions, though as indicated the effective
theory divides the spin response into separate longitudinal ($L_{JM}^5$) and transverse ($T_{JM}^{\mathrm{el}~5}$) components. These
spin responses have distinct
$R_k ( {v}_T^{ 2}, {q}^{\,2}/m_N^2 )$ -- different bilinear functions of the $c_i$ and thus different sensitivity to
the underlying EFT operators -- as well as $W_k(q^2b^2)$ characterized by distinct form factors.
In addition we find three new response functions, each proportional to $q/m_N$ (an
indication that they are connected with the finite site of the nucleus), generated in two cases by
familiar scalar and vector operators, $\vec{\sigma}(i) \cdot \vec{\ell}(i)$ and $\vec{\ell}(i)$, and in the third case by a
tensor operator that is clearly closely related to $\vec{\sigma}(i) \cdot \vec{\ell}(i)$. For one response function,
the vector longitudinal response $L_{JM}$, both scalar and tensor terms survive in the long wavelength limit.
One sees that the general form of the nuclear response, even in the long wavelength limit, is more
complex than is apparent from the SI/SD formulation. Two scalar ($J=0$) operators emerge from the effective theory,
and three vector ($J=1$) operators. Two sets of operators interfere, $C_{JM}-L_{JM}$ and $T_{JM}^{\mathrm{el}~5}-T_{JM}^\mathrm{mag}$, as happens in familiar electroweak processes such as neutrino scattering off nuclei.
The bottom line for experimentalists is that each nuclear response function is a ``knob" that
can be turned by picking nuclear targets with the requisite properties. Each response function is accompanied
by a WIMP tensor $R_k ( {v}_T^{2}, {q}^{2}/m_N^2 )$ that involves a distinct set of LECs.
Thus, including the two interference terms, in principle one can obtain eight constraints on the LECs
of the effective theory from elastic scattering experiments. In the future, when dark matter
scattering is observed, the experimental task will have been completed when those eight constraints
are determined: all information on candidate fundamental theories (the ultra-violet theories) that can be extracted
from inelastic scattering will have been obtained, at that point.
Experimentalists can turn the nuclear physics ``knobs" by selecting nuclear targets that isolate the various
response functions, thereby determining the associated $R_k ( {v}_T^{2}, {q}^{2}/m_N^2 )$. For example, if
one wants to extract the WIMP physics associated with the two response functions
generated by $\vec{\sigma}(i)$ and $\vec{\ell}(i)$, one could employ two odd-A nuclear targets,
one having its unpaired valence nucleon in an $\ell=0$ orbit
(an s-wave orbit), the other with a valence nucleon in a high-$\ell$ orbit. The former would be sensitive only to
$\vec{\sigma}(i)$, the latter primarily to $\vec{\ell}(i)$. In fact, one should do four such experiments
two with targets having unpaired valence protons and two with targets having unpaired valence neutrons, to test
the isospin of these couplings. With such a strategy, extended to the full set of nuclear response functions,
one would eventually learn everything than can
be learned from elastic scattering.
\section{The Point-Nucleus Limit: The Origin of the Additional Response Functions}
Before turning to the $R_k ( {v}_T^{2}, {q}^{2}/m_N^2 )$, it is interesting to explore the
origin of the additional response functions. The standard SI/SD form of the cross section can be obtained
(apart from the detail that one should allow the longitudinal and transverse projections of spin to
have distinct coefficients)
by taking the point nucleus limit. Naively one might hope that this point-nucleus SI/SD form corresponds
to some lowest-order approximation, capturing the leading-order effects of candidate effective interactions.
But in fact this is not the case for about half of the operators appearing in Eq. (\ref{eq:ops}), those involving the
WIMP-nucleon relative velocity.
As it is easiest to illustrate the point in an example, consider
\[ \mathcal{O}_8 = \sum_{i=1}^A \vec{S}_\chi(i) \cdot \vec{v}^\perp(i),~~~~~~~~~~\vec{v}^\perp(i) \equiv \vec{v}_\chi - \vec{v}_N(i). \]
(We have made a slight simplification by not symmetrizing the velocity between initial and final states --
see \cite{Liam2013} for details.) If we take the point-nucleus limit, all nucleons are at the same point, and the
interaction reduces to
\[ \vec{S}_\chi(i) \cdot \vec{v}_T \sum_{i=1}^A 1(i) ,\]
so that the nuclear interaction is SI with a coefficient that depends on the WIMP velocity relative to the
nuclear center-of-mass, $v_T \sim 10^{-3}$. Associated cross sections, which depend on the
square of this amplitude, would be of order $\sim 10^{-6}$ of the weak interaction value, consequently.
However it is easy to see that this point-nucleus limit is unjustified. The original interaction
contains $A$ independent internal WIMP-nucleon velocities, which we are free to rewrite by a
standard Jacobi transformation as
\[ \{\vec{v}^\perp(i), i=1,....,A \}~ \rightarrow~ \{ \vec{v}_T;~ \vec{\dot{v}}(i),i=1,....A-1 \} \]
singling out the WIMP velocity relative to the nuclear center-of-mass, leaving
$A-1$ velocities $\vec{\dot{v}}(i)$ corresponding to the independent internucleon velocities.
For example, for the simplest case of $A$=2, there is one such
internucleon velocity, $\vec{\dot{v}}_1 \equiv (\vec{v}(2)-\vec{v}(1))/\sqrt{2}$.
Now internucleon velocities are typically $\sim 10^{-1}$, not $\sim 10^{-3}$. Thus the point-nucleus limit
effectively keeps one amplitude that is tiny (proportional to $v_T$), and discards $A-1$ other velocities that couple to $\vec{S}_\chi$
with precisely the same strength, that are 100 times larger.
One might be mislead into thinking that one can neglect the internucleon velocities, despite their size,
because of parity: they are intrinsic nuclear operators carrying odd parity. But this overlooks the fact that
while the energy transfer in WIMP scattering is small, the three-momentum transfer is large, typically
$q r(i) \gtrsim 1$. Thus expanding the full nuclear operator to first order in q (while quantizing along
$\vec{q}$ and taking the $J=1$ part) yields
\[ e^{i \vec{q} \cdot \vec{r}(i)} \vec{\dot{v}}(i)~ \rightarrow~ -{1 \over i} q \vec{r}(i) \times \vec{\dot{v}}(i) =
-{1 \over i} {q \over m_N} \vec{r}(i) \times \vec{\dot{p}}(i) = -{q \over m_N} \vec{\ell}(i) \]
One obtains a familiar dimensionless parity-conserving operator, the orbital angular momentum,
accompanied by a dimensionless constant $q/m_N \sim 1/10 $. That is, our derivation shows that
the scale of the internal relative nucleon velocities is encoded in the parameter $q/m_N$.
This is then very helpful in the effective theory formulation. We mentioned previously that the most general Hermitian
WIMP-nucleon interaction can be constructed from the four dimensionless variables
\[ \vec{S}_\chi~~~~~\vec{S}_N~~~~~\vec{v}^\perp~~~~~i{\vec{q} \over m_N} \]
The reason that $m_N$ is the natural choice for the mass scale in the fourth variable is that the
parameter $q/m_N$ naturally arises from the proper treatment of relative nucleon velocities in the nucleus.
As outlined below (and discussed in greater detail in \cite{Nikhil2014}), when velocity-dependent
interactions are treated in a manner that properly takes into account nuclear
finite size, their effects are enhanced by a factor
of $\sim (\mu_T/m_N)^2 \sim 10^4$, where $\mu_T$ is the reduced mass for the WIMP-nucleus
scattering, relative to treatments that take the point-nucleus limit.
Thus velocity-dependent interactions that appear to be very difficult to test, in the point nucleus
limit, in fact make large contributions. These enhancement are connected with
new response functions generated by operators like $\vec{\ell}(i)$: by exploiting these new response functions,
one can determine the strengths of such velocity-dependent interactions. In addition to misrepresenting the
magnitude of the cross section,
the point nucleus limit also mischaracterizes the multipolarity of the scattering. This is seen in
our example, $\vec{S}_\chi \cdot \vec{v}^\perp$. The associated point-nucleus operator is $1(i)$,
a scalar, while the scattering is actually dominantly vector, generated by $\vec{\ell}(i)$.
\section{Constraining the Effective Theory Parameters}
The WIMP-nucleus differential cross section as a function of the recoil energy $E_R$ carried off by the
scattered nucleus can be written
\begin{equation}
{ d \sigma(v_T,E_R) \over dE_R} = {2 m_T \over \pi v_T^2} G_F^2 \left[ {1 \over 2 j_\chi+1}{1 \over 2j_N+1} \sum_\mathrm{spins} \left| \mathcal{M}^\mathrm{Nuc} \right|^2 \right]= {2 m_T \over \pi v_T^2} G_F^2 {4 \pi \over 2j_N+1}
\sum_{\tau,\tau^\prime=0}^1 \sum_k R_k^{\tau \tau^\prime} \left( v_T^{ 2}, {q^{\,2} \over m_N^2} \right)
~W^{\tau \tau^\prime}_k( q^{\,2} b^2)
\end{equation}
where the sum extends over the various response functions, generated by the six sets of multipole operators
including two interference terms. The Fermi constant has been introduced to make the weak scale explicit. The
sums over the isospin indices $\tau,\tau^\prime$ appear because the
$R_k^{\tau \tau^\prime}$ depend on the LECs
\[ c_i^{\tau=0} =(c_i^{(p)} + c_i^{(n)})/2~~~~~~~~~~ c_i^{\tau=1} =(c_i^{(p)} - c_i^{(n)})/2 \]
we introduced previously. Because we have normalized these to the weak scale,
the LECs are dimensionless. The conventions employed here
are those of Ref. \cite{Nikhil2014}, where more details are given.
In this formula, the left hand side will be measured (some day, hopefully), while on the right hand side, the experimentalist
can vary the $W^{\tau \tau^\prime}_k( q^{\,2} b^2)$, the nuclear response functions. These
response functions, as $q \rightarrow 0$, are generated by squaring the nuclear matrix elements
of the long-wavelength operators discussed in the
previous section; for general $q$, they
also involve nuclear form factors,
which in the case of the harmonic oscillator shell model, depend on $(qb)^2$, where $b$ is the harmonic oscillator size parameter. Consequently, in an ideal world, experimentalists would vary the $W^{\tau \tau^\prime}_k( q^{\,2} b^2)$
by using nuclear targets with different angular momenta J, different isospins, and different shell structures, to obtain
a sufficient number of constraints
to map out the eight $R_k^{\tau \tau^\prime} \left( v_T^{ 2}, q^{\,2}/m_N^2 \right)$ (six response functions
plus two interference terms). One would have then
obtained all of the information on dark matter particle interactions that is available from elastic scattering.
Note that more than eight constraints can be obtained: the measurements can be done as a function of $E_R =q^2/2 m_T$,
$m_T$ the target mass, to separate the LECs contributing to a given $R_k^{\tau \tau^\prime} \left( v_T^{ 2}, {q^{\,2} \over m_N^2} \right)$.
It is helpful to consider examples in which we contrast the standard SI/SD treatment with a full quantum mechanical
treatment, while still employing the EFT Lagrangian. For SI scattering we find
\begin{equation}
R_{SI}^{\tau \tau^\prime} = c_1^\tau c_1^{\tau^\prime} +{j_\chi(j_\chi+1) \over 3} \left[ {q^2 \over m_N^2} v_T^2 c_5^\tau c_5^{\tau^\prime}+v_T^2 c_8^\tau c_8^{\tau^\prime} + {q^2 \over m_N^2} c_{11}^\tau c_{11}^{\tau^\prime} \right]
~~~~~W^{\tau \tau^\prime}_{SI} \underset{q \rightarrow 0} {\longrightarrow} {1 \over 4 \pi} \langle j_N || \sum_{i=1}^A t^\tau(i) ||j_N \rangle \langle j_N ||\sum_{i=1}^A t^{\tau^\prime}(i) || j_N \rangle
\end{equation}
where $j_\chi$ is the WIMP spin. We see that the EFT operators $\mathcal{O}_5$ and $\mathcal{O}_8$
contribute to SI scattering, but their contributions are suppressed by $\sim 10^{-8}$ and $\sim 10^{-6}$ respectively.
($\mathcal{O}_8$ was the velocity-dependent operastor example we discussed in the previous section.) But if
the nuclear finite size is properly treated, one obtains a $T^\mathrm{mag}$ response function as well
\begin{equation}
R_{T^\mathrm{mag}}^{\tau \tau^\prime} = {j_\chi (j_\chi+1) \over 3}\left[ {q^2 \over m_N^2} c_5^\tau c_5^{\tau^\prime}+c_8^\tau c_8^{\tau^\prime} \right]~~~~~W^{\tau \tau^\prime}_{T^\mathrm{mag}} \underset{q \rightarrow 0} {\longrightarrow} {1 \over 24 \pi}
{q^2 \over m_N^2} \langle j_N || \sum_{i=1}^A \ell(i) ~t^\tau(i) ||j_N \rangle \langle j_N ||\sum_{i=1}^A \ell(i) ~ t^{\tau^\prime}(i) || j_N \rangle
\end{equation}
The two EFT operators $\mathcal{O}_5$ and $\mathcal{O}_8$ now are associated with kinematic suppression
factors of $10^{-4}$ and $10^{-2}$, respectively. Comparing the two results, we see that a proper treatment of
finite size
\begin{itemize}
\item allows us to separately measure the effects of $c_5$ and $c_8$, when previously
they appeared in combination with $c_1$ and $c_{11}$;
\item generates a response with increased sensitivity to
these two couplings; and
\item leads to a dominant rank-one response requiring $j_N>0$, in contrast to the point-nucleus result that
$\mathcal{O}_5$ and $\mathcal{O}_8$ contribute very weakly to the scalar SI response.
\end{itemize}
Similar in the SD point-nucleus limit
\begin{eqnarray}
R_{SD}^{\tau \tau^\prime}={1 \over 12} \left[ {q^2 \over m_N^2} v_T^2 c_3^\tau c_3^{\tau^\prime}+v_T^2 c_7^\tau c_7^{\tau^\prime}+{q^2 \over m_N^2} c_{10}^\tau c_{10}^{\tau^\prime} +{j_\chi(j_\chi+1) \over 3} \left( 3 c_4^\tau c_4^{\tau^\prime} +{q^2 \over m_N^2} (c_4^\tau c_6^{\tau^\prime}+c_6^\tau c_4^{\tau^\prime}
+2 c_9^\tau c_9^{\tau^\prime}) \right. \right. \nonumber \\ \left. \left. +{q^4 \over m_N^4} c_6^\tau c_6^{\tau^\prime}+2 v_T^2 c_{12}^\tau c_{12}^{\tau^\prime} +{q^2 \over m_N^2} v_T^2 (c_{13}^\tau c_{13}^{\tau^\prime} + c_{14}^\tau c_{14}^{\tau^\prime}-c_{12}^\tau c_{15}^{\tau^\prime} -c_{15}^\tau c_{12}^{\tau^\prime} ) + {q^4 \over m_N^4} v_T^2 c_{15}^\tau c_{15}^{\tau^\prime} \right) \right] \nonumber \\
W^{\tau \tau^\prime}_{SD} \underset{q \rightarrow 0} {\longrightarrow} {1 \over 4 \pi}
\langle j_N || \sum_{i=1}^A \sigma(i) ~t^\tau(i) ||j_N \rangle \langle j_N ||\sum_{i=1}^A \sigma(i) ~ t^{\tau^\prime}(i) || j_N \rangle~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
\end{eqnarray}
But the velocity-suppressed operators appearing above make much stronger contributions to, and can be extracted
more cleanly from, the new response functions that arise from properly treating the structure of the nucleus.
For example, for scattering off a $j_N=0$ nucleus, the longitudinal response function is
\begin{eqnarray}
R_L={q^2 \over4 m_N^2}c_3^\tau c_3^{\tau^\prime} + {j_\chi(j_\chi+1) \over 12} \left(c_{12}^\tau -{q^2 \over m_N^2} c_{15}^\tau \right) \left(
c_{12}^{\tau^\prime} - {q^2 \over m_N^2} c_{15}^{\tau^\prime} \right)~~~~~~~~~~~~~~~~~~~~ \nonumber \\
W^{\tau \tau^\prime}_{L} \underset{q \rightarrow 0} {\longrightarrow} {1 \over 36 \pi}
{q^2 \over m_N^2} \langle j_N || \sum_{i=1}^A \vec{\sigma}(i) \cdot \vec{ \ell}(i) ~t^\tau(i) ||j_N \rangle \langle j_N ||\sum_{i=1}^A \vec{\sigma}(i) \cdot \vec{\ell}(i) ~ t^{\tau^\prime}(i) || j_N \rangle
\end{eqnarray}
Here, as in our previous example, $\mathcal{O}_3$, $\mathcal{O}_{12}$, and $\mathcal{O}_{15}$ make substantially larger
contributions to a new scalar response than they do to the SD response. In a SI/SD treatment, one would
conclude that these operators appear in a complicated way
in the SD response, entwined with $\mathcal{O}_4$, $\mathcal{O}_6$, $\mathcal{O}_7$, $\mathcal{O}_{9}$, $\mathcal{O}_{10}$, $\mathcal{O}_{13}$, and $\mathcal{O}_{14}$ and suppressed by $v_T^2$,
when in fact they make the leading
contributions to a new type of scalar response. In principle all three couplings could be determined from the
nuclear recoil distribution, as
$\mathcal{O}_3$, $\mathcal{O}_{12}$, and $\mathcal{O}_{15}$ appear in $R_L$ with distinct leading-order behaviors in
$q^2/m_N^2$.
These few examples show that many more tools become available for constraining the effective theory's LECs once the
new response functions associated with nuclear compositeness are recognized. Velocity-dependent interactions
that appear to be difficult to probe in the standard SI/SD formalism can be isolated and studied in these
new response functions, where they play enhanced roles.
\section{Implications for Experiments}
In the previous section we have noted that the larger number of responses that are revealed when the
nuclear finite size is properly treated opens up lots of additional opportunities for constraining the
parameters of the effective theory, and thus for restricting the form of candidate ultraviolet theories
that map onto the EFT. Conversely, if one limits consideration to the standard SI/SD description
of the scattering, one can be mislead. For example, conflicts may appear to arise
between experiments because the analysis framework is too restrictive, when in fact no conflict exists.
\begin{figure}[t]
\includegraphics[width=150mm]{figure2.pdf}
\caption{A comparison of the spin response function (the longitudinal and transverse electric components of an
axial spin operator, left frames), employed in SI/SD analyses, with the orbital angular momentum response function
(the transverse magnetic component of the vector velocity operator, right frames). The upper (lower) frames assume a
coupling to protons (neutrons). The calculations are taken from the shell-model work of \cite{Liam2013,Nikhil2014}. }
\label{fig:fig2}
\end{figure}
One illustration in given in Fig. \ref{fig:fig2}, where the spin response function used in standard SI/SD
responses is compared to the transverse magnetic (orbital angular momentum) response function, for the targets F, Na, Ge, I, and Xe. The results
come from the shell-model work of \cite{Liam2013}. Both response functions are ``SD" in the sense that they
involve rank-one operators -- they cannot be distinguished geometrically -- but as they depend on distinct
underlying nuclear operators, they can be distinguished by selecting targets with the right properties.
In fact there are large differences in the responses to the
two underlying operators for targets currently in use in dark matter studies. For example, if the operators couple to protons,
the ratio in the responses of iodine (e.g., DAMA, KIMS, COUPP)
and fluorine (e.g., PICASSO, PICO, COUPP) for $\langle \vec{\sigma} \rangle^2$ vs. $\langle \vec{\ell} \rangle^2$
differ by $\sim$ 110. For couplings to neutrons, the responses of Xe (e.g., LUX, Xenon, XMASS) are similar for spin and orbital
angular momentum, but those for I and Ge (e.g., CDMS, COGENT, Edelweiss) change by an order of magnitude, while those for
Na (e.g., DAMA, ANAIS) and F change by two orders of magnitude. Consequently, if one analyzes experiments assuming
a SD coupling, but in fact the underlying operator is $\vec{\ell}(i)$, experiments that are
consistent may appear to be in conflict. Properly interpreted, however, one would find
instead evidence that the underlying WIMP-nucleon interaction is
associated with $\vec{\ell}(i)$, not $\vec{\sigma}(i)$: this ansatz could then be tested in a follow-up experiment.
The field is entering the ``G2" phase of dark matter elastic scattering experiments: a smaller number
of experiments may be done, but with larger detector masses. Our effective theory arguments, however,
argue for continued development of detectors using a variety of nuclear targets. If dark matter is seen
in direct-detection experiments, there is no doubt that a large set of experiments will be needed:
as established here, a great deal of information can be extracted from such experiments because of
the multiple response functions that contribute. Thus it is important, even if only a few detectors
are in operation, to continue developing new technologies that can be deployed when needed.
One technique and one target will not answer all of the question. Even early
in the discovery phase, it will be important to analyze experiments in the kind of general framework
provided here. Repeating an example from the paragraph above, if rates in a Ge detector appear to be high compared to those seen
in a Xe detector, this might suggest $\vec{\ell}(i)$ as an operator and thus influence the choice of a
third detector.
We would like to acknowledge useful conversations with Spencer Chang, Tim Cohen, Eugenio Del Nobile, Paddy Fox, Ami Katz, and Tim Tait. NA thanks the UC MultiCampus Research Program Dark Matter Search Initiative
for support. ALF was partially supported by ERC grant BSMOXFORD no. 228169. WH is supported by the US Department of Energy under contracts DE-SC00046548 and DE-AC02-98CH10886.
|
\section{Introduction}
Let $\pi$ be a permutation on $[n]$ and $\tau$ be a permutation on $[k].$ Then $\tau$ is said to occur as a {\em pattern} in $\pi$ if for some subsequence of $\pi$ of length $k$ all the elements of the subsequence occur in the same relative order as do the elements of $\tau.$ For example $1324$ occurs as a pattern in $152364$ as $1526$ and $1536$ as both are in the same relative order as $1324.$ If a permutation $\tau$ does not occur in $\pi,$ then this is said to be a {\em pattern-avoiding permutation,} or PAP.
Let $P(z) = \sum_{n \ge 0} p_nz^n$ be the ordinary generating function (OGF) for the number of permutations $p_n$ of length $n$ avoiding the pattern $1324.$
It is well known that, for the classical, length 4 PAPs, the 24 possible patterns fall into one of three possible classes \cite{ES10}, called Wilf classes. That is to say, there are three distinct OGFs describing all 24 patterns.
For the sequence $1234$ and its associated patterns, in 1990 Gessel \cite{IG90} showed that the number of length $n >0$ pattern-avoiding permutations is
\begin{equation} \label{eq:1234}
p_n(1234) = \frac{1}{(n+1)^2(n+2)}\sum_{k=0}^n \binom{2k}{k} \binom{n+1}{k+1} \binom{n+2}{k+1}.
\end{equation}
Asymptotically, $$p_n(1234) \sim 2.8 \cdot 9^n \cdot n^{-4},$$
and the generating function $P_{1234}(x) = \sum_n p_n(1234)x^n$ satisfies the linear ODE
\begin{multline}
(9x^5-19x^4+11x^3-x^2)\cdot \frac{d^3P_{1234}(x)}{dx^3} +(72x^4-153x^3+90x^2-9x)\cdot \frac{d^2P_{1234}(x)}{dx^2} + \\
+(126x^3-264x^2+154x-16)\cdot \frac{dP_{1234}(x)}{dx} + (32-72x+36x^2)\cdot P_{1234}(x)=0, \,\,\,\,\,\,\,\,\,\,\,\, \\
{\rm with\,\,\, initial\,\,\, conditions\,\,\,} P_{1234}(0)=1, \,\, P'_{1234}(2)=0, \,\, P''_{1234}(2)=12.\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
\end{multline}
For the sequence $1342$ and its associated patterns, in 1997 B\'ona \cite{MB97} showed that the number of length $n >0$ pattern-avoiding permutations is
\begin{equation} \label{eq:1342}
p_n(1342) = (-1)^{n-1} \cdot \frac{(7n^2-3n-2)}{2} + 3\sum_{k=0}^n (-1)^{n-i}\cdot 2^{i+1} \cdot \frac{(2i-4)!}{i!(i-2)!} \cdot\binom{n-i+2}{2}.
\end{equation}
The generating function $P_{1342}(x) = \sum_n p_n(1342)x^n$ satisfies the linear ODE
\begin{equation}
(8x^2+7x-1)\cdot \frac{d^2P_{1342}(x)}{dx^2}+(28x-8)\cdot \frac{dP_{1342}(x)}{dx} + 12\cdot P_{1342}(x)=0, \,\, P_{1342}(0)=1, \,\, P'_{1342}(0)=1.
\end{equation}
Indeed, it can be exactly solved to give \cite{K11} the simple algebraic expression
$$P_{1342}(x)=\frac{32x}{-8x^2+20x+1-(1-8x)^{3/2}},$$ from which one readily obtains $$p_n(1342) \sim \frac{64}{243 \cdot \sqrt{\pi}} \cdot 8^n \cdot n^{-5/2}.$$
The remaining class, that of 1324-avoiding permutations, remains unsolved \cite{ES12}. Even the growth constant is not accurately known. The best upper bound is 13.73718 \cite{MB14}, due to B\'ona, while Claesson, Jel\'inek and Steingr\'imsson \cite{CJS12} gave an improved bound, $e^{\pi \sqrt{2/3}} \approx 13.00195,$ but subject to the validity of an unproved conjecture. The best published lower bound is 9.47, proved by Albert et al. \cite{AR06} while David~Bevan has an as yet unpublished bound of 9.81. Careful Monte Carlo work by Madras and Liu \cite{ML10} implies that the growth constant lies in the range $[10.71,11.83].$
In this paper we give details of an improved algorithm for the enumeration of such PAPs, with which we obtained five further terms in the OGF beyond the existing longest known sequence, due to Johansson and Nakamura \cite{JN13}, using comparable computing resources. We will refer to their algorithm as the JN algorithm.
We also analyse the sequence of coefficients, and provide compelling numerical evidence that the asymptotic form of the coefficients is more complex than that of the two solved classes just considered. The generating functions of the two solved cases have algebraic singularities, whose coefficients have the asymptotic form $p_n \sim D \cdot \mu^n \cdot n^g.$ Our numerical studies, detailed below, lead us to suggest that the coefficients $p_n(1324) \sim B \cdot \mu^n \cdot \mu_1^{n^\sigma} \cdot n^g,$ where $\sigma = \frac{1}{2}.$
Many enumeration problems in algebraic combinatorics have generating functions with algebraic singularities, and hence coefficients with leading asymptotic form
\begin{equation} \label{eq:alg}
a_n \sim A \cdot \mu^n \cdot n^g,
\end{equation}
where $1/\mu$ is the radius of convergence, sometimes called the {\em critical point}, $g$ is a critical exponent and $A$ is a critical amplitude. However
a number of solved, and, we believe, unsolved problems that arise in both algebraic combinatorics and mathematical physics have a more complex singularity structure, with a sub-dominant asymptotic term ${\rm O}(\mu_1^{n^\sigma})$ rather then ${\rm O}(n^g).$ In fact the sub-sub dominant term is of ${\rm O}(n^g)$. That is to say, the dominant asymptotic behaviour of the coefficients $b_n$ of the associated generating function is
\begin{equation} \label{eq:can1}
b_n \sim B \cdot \mu^n \cdot \mu_1^{n^\sigma} \cdot n^{g}.
\end{equation}
Perhaps the best-known example of this sort of behaviour is the number of partitions of the integers -- though in that case the leading exponential growth term $\mu^n$ is absent (or, equivalently, $\mu=1$). Another example is the generating function for the number of fragmented permutations \cite{FS09}, which is $$F(z) =\exp\left ( \frac{z}{1-z} \right ).$$ Then, with $F_n=[z^n]F(z),$ we have\cite{FS09}, p563, $$F_n \sim \frac{e^{2\sqrt{n}}}{2\sqrt{\pi e}n^{3/4}}.$$ These two examples highlight the fact that singularities of quite different analytic structure can give rise to the same asymptotics. The OGF for integer partitions which has radius of convergence $r_c=1,$ has a natural boundary on the unit circle, whereas the generating function for fragmented permutations is D-finite. So if such an asymptotic form (\ref{eq:can1}) is observed, one cannot say much about the underlying singularity structure from the asymptotics alone.
Another example, discussed in \cite{G14}, is the problem of Dyck paths counted not only by length but also by height $h$, which is defined to be the maximal vertical displacement of the Dyck path from its horizontal axis. Let $d_{n,h} $ be the number of Dyck paths of length $2n$ and height $h.$ The generating function is
$$D(x,y)=\sum_{n,h}d_{n,h}x^{2n}y^h.$$
Then
\begin{equation} \label{eq:Dyck}
[x^{2n}]D(x,y)=\sum_{h=1}^n d_{n,h}y^h \sim B\cdot 4^n \cdot \mu_1^{n^{1/3}} \cdot n^{-5/6},
\end{equation}
where both $B(y)$ and $\mu_1(y)$ are known \cite{G14}.
There are also a number of models in the mathematical physics literature that have this more complex asymptotic structure. In particular, Duplantier and Saleur \cite{DS87} and Duplantier and David \cite{DD88} studied the case of {\em dense} polymers in two dimensions, and found the partition functions had the asymptotic form (\ref{eq:can1}). In \cite{OPB93}, Owczarek, Prellberg and Brak investigated an exactly solvable model of interacting partially-directed self-avoiding walks (IPDSAW), for which the solution had previously been given by Brak, Guttmann and Whittington in \cite{BGW92}. In \cite{OPB93}, Owczarek et al. analysed a 6000 term series expansion for IPDSAWs in the collapsed regime, and estimated $\sigma = 1/2,$ $g = -3/4,$ while $\mu_1$ was estimated to at least 6 digit accuracy. From \cite{BGW92} the value of $\mu$ is exactly known. Subsequently Duplantier \cite{D93} pointed out that $\sigma = 1/2$ is to be expected, not only for IPDSAWs, but also for SAWs in the collapsed regime, for the two-dimensional version of these models. In all the examples we have encountered, $\sigma$ takes the value 1/3, 1/2 or 2/3.
In the next section we give details of the enumeration algorithm. In subsequent sections we analyse the available series coefficients.
\section{Algorithm}
The algorithm used can be considered to be a set of further optimizations on the JN algorithm.
However a significantly different notation is used as this helps make some of the optimizations
clearer, as well as helping with a memory efficient encoding implementation. We note that Marinov and Rodoi\v{c}i\'c \cite{MR03} have previously given a recursive algorithm for this problem. It is a significantly different algorithm conceptually to the JN algorithm, although one could imagine a variant on that algorithm keeping track of 12 patterns as part of their labels, instead of their set of $l(\pi)$ values. Such an algorithm would probably be similar in performance to this algorithm.
\subsection{Basic Algorithm}
We will start with a very simple (and inefficient) algorithm. Let $f(n,P)$ be the number
of permutations avoiding the pattern 1324 with $n$ numbers remaining and starting with
the prefix $P$ (a sequence of integers). Then the desired series is $f(n,\varnothing)$. Each
value $f(n,P)$ can be expressed as the sum of up to $n$ other terms $f(n-1,P')$ where $P'$ is
$P$ followed by one extra integer. There will be fewer than $n$ terms if $P'$ implies a
1324 pattern, usually due to containing a 132 pattern with a 4 inevitably to eventually follow.
Using the termination condition $f(0,-)=1$, one could easily
encode a recursive algorithm that would work.
This algorithm will systematically individually enumerate every permutation avoiding 1324,
and its time consumption will be proportional to the answer. Like many such recursive
enumeration algorithms, one can get a much faster algorithm by recognising that there
exist many classes $S = (n,{P_1,P_2,P_3,...})$ such that $f(n,P_i)=f(n,P_j)$ for all $i$ and $j$.
Define $f(S)$=$f(n,P_-), $ $P_-$ means any prefix from the class. Now, modify the algorithm to use $S$ (which will henceforth
be called a signature) in a recursive function $f(S)$. After computing
a value of $f(S)$, store it in some table. When you next need $f(S)$, look it up in the
table. If it is already there, then use the stored value. This can vastly improve
the speed of execution as you will avoid passing through large swathes of the
enumeration tree. It does have a cost of memory. This approach is often called dynamic
programming, memoization, or memorization.
When using such approaches, the definition of the signatures $S$ is paramount. The
more prefixes you can prove to be identical (and thus members of the same signature),
the more efficient the algorithm will be. Indeed, the time and memory will both
be proportional to the number of different signatures. The rest of this subsection deals
with the definition of prefixes.
The signature must contain enough information to allow the algorithm to avoid
1324 patterns. One way to do this is to keep track of all the 13 patterns. Then,
in the recurrence relation, you do not allow any numbers in the middle of
one of these 13 patterns (i.e. a 2) if there are any numbers remaining bigger than
the 3 (i.e. a 4, which would then inevitably follow at some point).
There is no need to keep track of exactly what the numbers are; you just need
to know the total number of numbers between, above, and below each 13 pattern.
One suitable notation to keep track of the 13 patterns is as a partition of
integers (the number of numbers left to go) with a well formed set of brackets
(the 13 patterns). It is also necessary to keep track of the lowest number
so far in the prefix, as all future numbers higher than it will form a 13
pattern with it. They may of course form other 13 patterns, but the one with the
lowest 1 will be the most restrictive, and the others may be ignored. The lowest
number is recorded with a comma. For brevity, the comma may be left out if there is a bracket
immediately following it.
Some typical signatures are shown in table \ref{tab:egSigs}, for some example prefixes and initial $n=20$.
\begin{table}
\caption{Example prefixes and corresponding signatures for $n=20$.}
\label{tab:egSigs}
\begin{center}
\begin{tabular}{lll} \hline \hline
Prefix & Signature & Explanation \\
\hline \hline
$\varnothing$ & 20 & 20 numbers left to go \\
11 & 10,9 & ten numbers to the left of the lowest number, 9 to the right \\
11,14 & 10,[2]6 & a 13 pattern introduced with two numbers between it, and 6 to the right \\
11,14,5 & 4,5[2]6 & a new lowest number, but no new 13 pairs \\
11,14,5,9 & 4,[3]1[2]6 & a new 13 pair produced \\
11,14,5,9,15 & 4,[[3]1[2]]5 & a new 13 pair produced \\
\hline \hline
\end{tabular}
\end{center}
\end{table}
At this point one can simplify the signature. A signature of the form $a[b[c]]d$ (that is,
with two consecutive closing brackets) has two restrictions.
The outer bracket means that you can't have anything in $b$ or $c$
until $d$ is finished. The inner bracket means that you can't have anything in $c$ until
$d$ is finished. The outer bracket is strictly more restrictive, so the inner one
is redundant and may be removed. So $a[b[c]]d=a[bc]d$. This simplification means that we will never have
two closing brackets in succession.
Adjacent integers not separated by brackets or a comma can be added together. For instance, $2[3[4]]5$, after removing the inner bracket, becomes
$2[7]5$ rather than needing to record that the inner $7$ was at one point broken unto a $3$ and a $4$.
Any tail bracket at the end of the signature of course can be removed; the brackets
are only restrictive if there are larger numbers possible.
At this point there is an isomorphism to the JN algorithm \cite{JN13}. Their final functional
form (last equation in section 2) is $H^0_n (t; b_1, ... , b_n; k)$. Here $k$ encodes the
position of the comma, and $b_i$ encodes the location of the open bracket corresponding
to a closing bracket at position $i$. Enumerating using these signatures produces an algorithm
basically identical in performance to the JN algorithm. The further simplifications described
below will improve performance.
Repeated open brackets can also be simplified. Consider a signature of the form $a[[b]c]d$. The
outer bracket means you cannot have anything in $b$ or $c$ until everything in $d$ is done.
The inner bracket means you cannot have anything in $b$ until $c$ is done. This is equivalent
to the restrictions described by the signature $a[b][c]d$. Simplifying signatures
by getting rid of all consecutive open brackets reduces the total number of signatures
significantly, in practice by a factor of roughly 4.
A minor simplification comes from dealing with open brackets at the start of the
signature. $[a]b$ means that everything in $b$ must be dealt with before everything in $a$.
This means that they are decoupled; indeed $f([a]b) = f(a)f(b)$. Factorizing the problem
seems like a big advantage, but only a small proportion of signatures start with brackets;
in practice this reduces the number of signatures by a factor of roughly 2.
An example of all the computations done (in the order that they are finished)
for permutations of length 6 is given in table \ref{tab:exhaustiveList}
\begin{table}
\caption{ All signatures used to compute up to the $n=6$ term. Single lines are used when a higher $n$ is started. }
\label{tab:exhaustiveList}
\begin{center}
\begin{tabular}{lll} \hline \hline
Signature $S$ & Composed of & $f(S)$ \\
\hline \hline
1 & , & 1 \\
\hline
,1 & , & 1 \\
2 & ,1 + 1 & 2 \\
\hline
,2 & ,1 + ,1 & 2 \\
1,1 & ,1 + 1 & 2 \\
3 & ,2 + 1,1 + 2 & 6 \\
\hline
,3 & ,2 + [1]1 + ,2 & 5 \\
1,2 & ,2 + 1,1 + 1,1 & 6 \\
2,1 & ,2 + 1,1 + 2 & 6 \\
4 & ,3 + 1,2 + 2,1 + 3 & 23 \\
\hline
,4 & ,3 + [1]2 + [2]1 + ,3 & 14 \\
1[1]1 & [1]1 + 1,1 & 3 \\
1,3 & ,3 + 1,2 + 1[1]1 + 1,2 & 20 \\
2,2 & ,3 + 1,2 + 2,1 + 2,1 & 23 \\
3,1 & ,3 + 1,2 + 2,1 + 3 & 23 \\
5 & ,4 + 1,3 + 2,2 + 3,1 + 4 & 103 \\
\hline
,5 & ,4 + [1]3 + [2]2 + [3]1 + ,4 & 42 \\
1[1]2 & [1]2 + 1[1]1 + 1[1]1 & 8 \\
1[2]1 & [2]1 + 1,2 & 8 \\
1,4 & ,4 + 1,3 + 1[1]2 + 1[2]1 + 1,3 & 70 \\
,1[1]1 & [1]1 + ,2 & 3 \\
2[1]1 & ,1[1]1 + 1[1]1 + 2,1 & 12 \\
2,3 & ,4 + 1,3 + 2,2 + 2[1]1 + 2,2 & 92 \\
3,2 & ,4 + 1,3 + 2,2 + 3,1 + 3,1 & 103 \\
4,1 & ,4 + 1,3 + 2,2 + 3,1 + 4 & 103 \\
6 & ,5 + 1,4 + 2,3 + 3,2 + 4,1 + 5 & 513 \\
\hline \hline
\end{tabular}
\end{center}
\end{table}
\subsection{Other techniques to reduce memory consumption}
A more complex simplification comes from noticing that a signature of the form
$a[b]c$ will not touch any of the $b$ values until all of $c$ is dealt with. This
means that there will be a set of signature prefixes $p_i$ with multiplicities $m_i,$ independent of $b,$ such that
$f(a[b]c)$ = $\sum\limits_{i} m_i f(p_i b)$. These $p_i$ and $m_i$ can be computed when needed and
cached. If you did this for all $a[b]c$ you would spend more time and memory
on this optimization than the original problem involved. However, if you just
do it for sufficiently short $a$ (in practice we used length of $a < 8$) then it can
reduce memory consumption by about 30 percent without significant effect on speed.
Note that the length of $a$ is more important
than the length of $c$, as all prefixes $p_i$ must be no longer than $a$, so a small length
of $a$ ensures that the number of terms here does not get too large. Also note that
$[b]$ can actually be a series of brackets, e.g. $[b_1][b_2][b_3][b_4]$.
Reducing the number of signatures reduces both the memory and the execution time, but
memory consumption tends to be the bottleneck. One simple trick for problems of this
class is to only save the result some fraction $p$ of the time. Signatures that are
only used once may not take up memory, and frequently used signatures will get
stored eventually. The smaller $p$, the less memory used, but the more time used. We
found that $p=0.3$ reduced memory consumption by about 30 percent with about a 30 percent
increase in execution time. This indicates that a significant fraction of the values $f(S)$
computed are only used once.
\subsection{Implementation}
The algorithm was implemented in Scala, which compiles to Java virtual machine bytecode,
and run on a computer with 1TB RAM.
The signature described here is less straightforward to encode on a computer than
the array of 32 bytes in \cite{JN13}. However, in practice it can be easily done in 128 bits
for all the signatures needed for $n$ up to the 50s.
This makes the keys 16 bytes, reducing memory consumption.
The signatures are encoded as a bitstring. The bitstring starts with a 6 bit number
equal to the sum of the integers in the signature $(n).$ This is used for determining
when the bitstring stops. There is then one bit to indicate whether the signature starts
with a comma. Then there is a repeating series of 1 bit for whether there is an
open bracket, 2 to 8 bits for encoding an integer, and one bit
for whether there is a closing bracket after that integer. This repeats until everything
is encoded.
Integers were encoded as follows:
\begin{itemize}
\item 00 means the number 1,
\item 01 means the number 2,
\item 10bbb means the number bbb+3,
\item 11bbbbbb means the number bbbbbb+11
\end{itemize}
Some signatures have a large number of small numbers in them; this will encode them in
a small number of bits. Others have a small number of large numbers; this will also encode
them in a small number of bits. Indeed, up until $n=29$, only 64 bits are needed for
the keys.
The values are stored as 64 bit values {\em modulo} some large prime. The computation is redone
{\em modulo} a different prime and the values reassembled using the Chinese remainder theorem.
This somewhat reduces memory use relative to 128 bit values, but also simplifies implementation
as the java virtual machine does not handle 128 bit integers easily.
The main memoization store was therefore effectively a large 128 bit key to 64 bit value
hash table. As the upper 64 bits of the key were sparsely used (indeed only 131 different
values for 35 terms), a significant memory saving was generated by having a master hash map
from the upper 64 bits to a slave hash map. The slave hash map then only needed to use the
lower 64 bits as keys. In practice the slave hash maps were still further subdivided in
one more layer; this was due to java virtual machine limitations on array size restricting
the length of a hash map. But the end result is the main memory store used 64 bit keys
and values in a gnu trove hash map.
Various of the tricks used increased computation time to a matter of days; we made a
parallelized version which worked about twice as fast as the single threaded version
on a 4 core desktop pc for testing, but turned out slower on the 32 core production
machine than the single threaded version, possibly because of memory coherency overheads.
So it was not used.
The parallelization comes from having different threads compute different signatures.
If, in computing $f(S)$, the value $f(s)$ for a subsignature $s$ is not available, then
a placeholder is inserted in $f(S)$, and $s$ is added to a priority queue of signatures
to evaluate. When a processor has nothing else to do, it takes the signature with the
lowest $n$ from the priority queue and evaluates it. When a signature is finished, it
fills in all the placeholders for that signature. Taking the signature with the
lowest $n$ prevents the queue from growing exponentially with $n$.
With all these improvements implemented, the algorithm produced 5 further terms on a comparable computer. The calculation had to be run twice, {\em modulo} two different primes, and the result reconstructed by the Chinese Remainder Theorem. As a check, we also ran it a third time, {\em modulo} another prime.
For each of the two moduli used, the program ran for 5 days and used somewhat over 540 GB of memory. This is the amount of memory used at the termination of the algorithm as reported by the JVM; some extra is needed for various tasks, primarily temporary objects such as resizing of hash tables.
Source code is available on https://github.com/AndrewConway/enumeration/ in the folder
avoid1324.
The coefficients are given in Table \ref{tab:coeffs}.
\begin{table}[htbp]
\centering
\caption{Coefficients of $1324$ pattern-avoiding permutations.}
\label{tab:coeffs}
\begin{center}
\begin{tabular}{ll} \hline \hline
1& 3421888118907\\
1&25887131596018\\
2&198244731603623\\
6&1535346218316422\\
23&12015325816028313\\
103&94944352095728825\\
513&757046484552152932\\
2762&6087537591051072864\\
15793&49339914891701589053\\
94776&402890652358573525928\\
591950&3313004165660965754922\\
3824112&27424185239545986820514\\
25431452&228437994561962363104048\\
173453058&1914189093351633702834757\\
1209639642&16130725510342551986540152\\
8604450011&136664757387536091240503406\\
62300851632&1163812341034817216384582333\\
458374397312&9959364766841851088593974979\\
&85626551244475524038311935717\\
\hline
\end{tabular}
\end{center}
\end{table}
\section{Analysis}
In the case of a simple algebraic singularity with asymptotic form (\ref{eq:alg}), the ratio of the coefficients is
\begin{equation} \label{eq:rat1}
r_n = \frac{a_n}{a_{n-1}} \sim \mu\left (1 + \frac{g}{n} +{\rm O}(\frac{1}{n^2}) \right ).
\end{equation}
If on the other hand the coefficients of some generating function are as in eqn. (\ref{eq:can1}), then the ratio of successive coefficients $r_n = b_n/b_{n-1},$ is
\begin{multline} \label{eq:rn}
r_n = \mu \left (1 + \frac{\sigma \log \mu_1}{n^{1-\sigma}} + \frac{g}{n} + \frac{\sigma^2 \log^2 \mu_1}{2n^{2-2\sigma}} + \frac {(\sigma-\sigma^2)\log \mu_1+2g\sigma \log \mu_1}{2n^{2-\sigma}} \right . \\
\left . {}+ \frac{\sigma^3 \log^3 \mu_1}{6n^{3-3\sigma}} +{\rm O}(n^{2\sigma-3}) + {\rm O}(n^{-2}) \right ).
\end{multline}
In particular,
when $\sigma = \frac{1}{2},$ this specialises to
\begin{equation} \label{eq:half}
r_n = \mu \left (1 + \frac{ \log \mu_1}{2\sqrt{n}} + \frac{g+\frac{1}{8}\log^2 \mu_1}{n} + \frac{\log^3\mu_1+(6+24g)\log \mu_1 }{48n^{3/2} } + {\rm O}(n^{-2}) \right ).
\end{equation}
\subsection{Differential approximant analysis}
The most successful numerical method for extracting the asymptotics from the first few terms of the OGF of a function with an algebraic singularity\footnote{With a slight abuse of notation, we refer to a singularity of the form $(1-x/x_c)^{\alpha}$ as an algebraic singularity even in those cases where $\alpha$ is not rational.} is the method of differential approximants, (called the DA method) due to Guttmann and Joyce \cite{GJ72}, with subsequent refinements due to Baker and Hunter \cite{HB79} and Fisher and Au-Yang \cite{FA79}. Details are given in \cite{G14, GJ09, G89}. In brief, the method fits available coefficients to a judiciously chosen family of D-finite ordinary differential equations (ODEs), and the singularity structure of the ODEs is extracted by standard methods \cite{Ince27, Forsyth02}.
For models with an isolated algebraic singularity, the method is very successful, with the radius of convergence and critical exponents frequently estimated to 10 significant digit accuracy or better, from a series of length 30-80 terms. However, when the method is used to analyse models with singularities that are not algebraic, such as those whose coefficients have the asymptotic form (\ref{eq:can1}), the method fails, though in a predictable manner. That is to say, one finds that the radius of convergence estimates are typically only found to two or three significant digits, and the critical exponent estimates are numerically large, typically around 10 or -10.
In this way, the method is useful -- as is a canary in a coal mine. If one analyses the known terms of the series with the method of differential approximants and finds estimates of the radius of convergence to be poorly converged, with numerically large exponent values, one can be confident that the underlying OGF does not have an algebraic singularity. Applying the method to the first 30 terms of the 1342 and 1234 PAPs, the known solutions are found. Applying the method to the 36 coefficients we have for 1324 PAPs, the method suggests that the radius of convergence is around 0.09, with an exponent variously estimated to be -20 or +15 or anything in between! This is the hallmark of a non-algebraic singularity.
Further details of the DA method, its successes and limitations are discussed in \cite{G14}. For the moment, we simply conclude that the OGF of 1324 PAPs almost certainly does not have an algebraic singularity. In the next section we explore the nature of the singularity by looking at the ratio of successive coefficients.
\subsection{Ratio analysis}
In order to determine the nature of the asymptotic form of the coefficients of the 1324-PAP OGF,
we first plot the ratios of successive coefficients $r_n = p_n/p_{n-1}$ against $1/n,$ as shown in figure \ref{fig:ratios}(a). The locus is clearly concave. This is inconsistent with an algebraic singularity, as can be seen from eqn. (\ref{eq:rat1}). We next plot the same ratios against $1/\sqrt{n}$ in figure \ref{fig:ratios}(b), this time as a point plot, and the plot is seen to be visually linear, implying, from eqn (\ref{eq:rn}) that $\sigma \approx 1/2.$ The outlying point near to the vertical axis is a Monte Carlo result obtained by Steingr\'imsson \cite{ES12} for PAPs of length 1001.
Linear extrapolation (not including the isolated point) implies a limiting value as $n \to \infty$ around 11.5. We can significantly improve on this estimate by considering the sequence of extrapolants defined by successive pairs of points. That is to say, one can simply linearly extrapolate successive pairs of ratios $(r_k,r_{k+1})$ with $k$ increasing up to the maximum value achievable with our data, which is 35. A plot of successive extrapolants against $1/n$ is shown in figure \ref{fig:sig1}(a), which appears to be linear. A crude extrapolation with a ruler suggests a limit of around 11.60.
Assuming (tentatively) that $\sigma = 1/2,$ another way to use the ratios to get a better estimate of $\mu$ is to eliminate the assumed term O$(1/\sqrt{n})$ in the ratios by forming the modified ratios
\begin{equation} \label{eq:modrat}
intercept_n = \frac{\sqrt{n} \cdot r_n - \sqrt{n-1} \cdot r_{n-1}}{\sqrt{n} -\sqrt{n-1}} \sim \mu \cdot \left ( 1 + {\rm O} (1/n) \right).
\end{equation}
We show in Figure \ref{fig:sig1}(b) a plot of $intercept_n$ against $1/n,$ and this too appears to be going to a value very close to 11.60.
We will take that as our initial estimate, which we subsequently refine. We also take $\sigma =1/2$ as our (initial) conjectured value. In doing so we are, in part, relying on the observation that in all known cases \cite{G14} where this asymptotic behaviour is observed, $\sigma$ is a simple rational, usually $1/2$ or $1/3$.
We also plotted (but don't display) the ratios against $1/n^{2/3},$ which is appropriate if $\sigma = 1/3.$ In that case the locus was convex, rather than concave. So on the basis of ratio plots alone, $\sigma \approx 1/2$ is our tentative estimate.
\begin{figure}[t!]
\centering
\subfigure[Plot of ratios of coefficients against $\frac{1}{n}$.]{\includegraphics[width=7.3cm]{papr1.jpg}}
\subfigure[Plot of ratios of coefficients against $\frac{1}{\sqrt{n}}$.]{\includegraphics[width=7.3cm]{papr2.jpg}}
\caption{}
\label{fig:ratios}
\end{figure}
In order to more accurately determine the value of the exponent $\sigma$, we note from (\ref{eq:rn}) that $$(r_n/\mu-1) \sim const. n^{\sigma-1}.$$ We show in figure \ref{fig:sig2}(a) a log-log plot of $(1-r_n/\mu)$ against ${n},$ where we have taken 11.60 as the (tentative) value of $\mu.$ This should be linear, with gradient $\sigma-1.$ A small degree of curvature is evident. Accordingly, we extrapolate the local ratios, defined as
\begin{equation} \label{eq:locrat}
1-\sigma_n = \frac{\log \left ( 1 -\frac{r_{n-1}}{\mu} \right )-\log \left ( 1 -\frac{r_{n}}{\mu} \right )}{\log {n} -\log(n-1)},
\end{equation}
against $1/n.$ The results are shown in Figure \ref{fig:sig2}(b)
The ordinates are estimators of $1-\sigma.$ If one accepts that $\sigma$ is a simple rational number, the value $1/2$ is inescapable.
We can also estimate $\sigma$ without assuming the value of $\mu$ as follows. From eqn. (\ref{eq:can1}), one sees that
\begin{equation} \label{eq:sig1}
r_{\sigma_n} = \frac{b_n \cdot b_{n-2}}{b_{n-1}^2} \sim 1 + \frac{\sigma \cdot (\sigma-1)\log{\mu_1}}{n^{2-\sigma}} + {\rm O}(1/n^2),
\end{equation}
so $\sigma$ can be estimated from a log-log plot of $r_{\sigma_n}-1$ against $n,$ independent of the value of $\mu.$ From this plot, shown in Figure \ref{fig:sig3}(a), which is linear as expected, we calculate the local gradient at each pair of successive points, as described above, and plot these against $1/n.$
This is shown in figure \ref{fig:sig3}(b), and it can be seen that the expected limit, as $n \to 0$ is, plausibly, $-1.5.$ From eqn. (\ref{eq:sig1}), this limit should be $2 - \sigma,$ which is consistent with our assertion that $\sigma = 1/2.$
In our subsequent analysis, we will assume this value.
\begin{figure}[t!]
\centering
\subfigure[Plot of extrapolated ratios against $\frac{1}{n}$.]{\includegraphics[width=7.3cm]{papextrap.jpg}}
\subfigure[Plot of estimators of $\mu$ by square root intercepts.]{\includegraphics[width=7.3cm]{linint.jpg}}
\caption{}
\label{fig:sig1}
\end{figure}
\begin{figure}[t!]
\centering
\subfigure[Log-log plot of $(1-r_n/\mu)$ against $n$..]{\includegraphics[width=7.3cm]{pap11.jpg}}
\subfigure[Plot of local extrapolants of figure at left, estimating $1-\sigma.$]{\includegraphics[width=7.3cm]{sig1.jpg}}
\caption{}
\label{fig:sig2}
\end{figure}
\begin{figure}[t!]
\centering
\subfigure[Log-log plot of $r_{\sigma_n}$ against $n$.]{\includegraphics[width=7.3cm]{rsigll.jpg}}
\subfigure[Local gradient of log-log plots of $r_{\sigma_n}$ against $\frac{1}{n}$.]{\includegraphics[width=7.3cm]{1324-ratio-ratio.jpg}}
\caption{}
\label{fig:sig3}
\end{figure}
With $\sigma$ taken to be $1/2,$ we attempt to refine the estimate of $\mu$ by extrapolating the ratios of the coefficients $r_n$ using the Bulirsch-Stoer \cite{BS64} algorithm, with parameter $w=1/2.$ This algorithm extrapolates a sequence $\{s_n\}$ assuming $s_n \sim s_\infty + c/n^w + o(n^{-w}),$ where $w$ is provided by the user. The results are given in Table \ref{tab:BS3}. Each successive row represents a higher order of extrapolation. We only show the last 7 entries of each order of extrapolation. We continue generating rows until the row entries lose monotonicity. From the first row we conclude $\mu < 12.832.$ From the second we conclude $\mu < 11.6663,$ and from the third we conclude $\mu < 11.6112.$ The rate of decrease in the entries in the third row is consistent with our initial estimate of 11.60, so we will retain this estimate for the time being, as the extrapolation has not given us a more precise value. It does however add support to that choice.
\begin{table}
\caption{\label{tab:BS3}
Last seven entries in each row of the table of Bulirsch-Stoer extrapolants. Each successive row is the result of a successively higher degree of extrapolation. The ratios of successive coefficients are extrapolated here, with parameter $w=1/2.$}
\begin{center}
\begin{tabular}{llllllll} \hline \hline
L & T(L,N-L-6) & T(L,N-L-5) & T(L,N-L-4)& T(L,N-L-3) & T(L,N-L-2) & T(L,N-L-1) & T(L,N-L) \\
\hline
\hline
1&13.17792074& 13.10768870 &13.04333009 &12.98413360 &12.92949766 &12.87891071& 12.83193533 \\
2&11.67944917 &11.67746701 &11.67538180 &11.67320098 &11.67094284 &11.66863183 &11.66629498 \\
3&11.64370166 &11.63846233 &11.63275886 &11.62693780 &11.62127955 &11.61599904 &11.61124940 \\
4&11.73026868 &11.71850728 &11.71746270 &11.72167941 &11.73128989 &11.74947827 &11.78630320\\
\hline \hline
\end{tabular}
\end{center}
\end{table}
Assuming then that $\sigma=1/2,$ from (\ref{eq:half}), it follows that $$r_n/\mu = 1 + \frac{ \log \mu_1}{2\sqrt{n}} + \frac{g+\frac{1}{8}\log^2 \mu_1}{n} + {\rm O}(n^{-3/2}). $$
In order to estimate $\mu_1$ and $g,$ we solve, sequentially, the trio of equations
\begin{equation} \label{eq:fit2a}
r_j/\mu = 1 + \frac{ c_1}{\sqrt{j}} + \frac{c_2}{j} + \frac{c_3}{j^{3/2}},
\end{equation}
for $ j=k-1,$ $j=k$ and $j=k+1,$ with $k$ ranging from 2 up to 35, and $\mu$ set at 11.60.
The results are shown in figures \ref{fig:papc}(a) and \ref{fig:papc}(b), plotting the parameters $c_1$ and $c_2$ respectively. The first neglected term in the asymptotics is O$(1/n^2)$ which is O$(1/n^{3/2})$ smaller than the term with coefficient $c_1,$ so $c_1$ is plotted against $1/n^{3/2}.$ By a similar argument, $c_2$ is plotted against $1/{n}.$ A simple visual extrapolation gives the estimate $c_1 = -1.615 \pm 0.005.$ The plot for $c_2$ is difficult to extrapolate. It appears to be turning near its end point, and we very tentatively estimate $c_2 \approx 0.15.$ Unless the gradient changes sign, we can only say that $c_2 < 0.2,$ and it seems to be going to a positive value. If the gradient changes sign, we can't even say that. We don't show the plot of $c_3$ as we cannot extrapolate it. From (\ref{eq:half}), $c_1=\log{\mu_1}/2$ and $c_2=g+\frac{1}{8}\log^2 \mu_1.$ Hence we estimate $\log{\mu_1} \approx -3.23,$ and assuming $c_2$ is in the range $[0,0.2]$ this gives $g = -1.2 \pm 0.15.$ We repeated this analysis varying the estimate of $\mu$ in the range [10.58,11.62]. With $\mu$ in this range, we estimate $\log{\mu_1} = 3.23 \pm 0.07,$ and $g = -1.2 \pm 0.4.$
\begin{figure}[t!]
\centering
\subfigure[Plot of parameter $c_1$ of (\ref{eq:fit2a}) against $\frac{1}{n^{3/2}}$.]{\includegraphics[width=7.3cm]{papc1.jpg}}
\subfigure[Plot of parameter $c_2$ of (\ref{eq:fit2a}) against $\frac{1}{n}$.]{\includegraphics[width=7.3cm]{papc2.jpg}}
\caption{}
\label{fig:papc}
\end{figure}
An alternative form of analysis involves direct fitting to the parameters in the assumed asymptotic form. That is to say, the assumed asymptotic form is $$b_n \sim B\cdot \mu^n \cdot \mu_1^{n^\sigma} \cdot n^g.$$ Therefore
\begin{equation} \label{logcan1}
\log {b_n} \sim \log{B} + n \log{\mu} + n^\sigma \log{\mu_1} + g \log {n}.
\end{equation}
So if $\sigma$ is known, or assumed, we have four unknowns in this linear equation. It is then straightforward to solve the linear system
$$\log {b_k} = c_1k + c_2 k^\sigma +c_3 \log {k}+c_4$$ for $k=n-2,\, n-1, \, n, \, n+1$ with $n$ ranging from $3$ to $35.$ Then $c_1$ estimates $\log {\mu},$ $c_2$ estimates $\log{\mu_1}$, $c_3$ estimates $g$ and $c_4$ gives estimators of $\log{B}.$
An obvious useful variation is in those cases where, say, $\mu$ is known, or accurately estimated. Then one can solve
\begin{equation} \label{eq:three}
\log b_n - \mu \log{n} = c_1 n^\sigma +c_2 \log {n} +c_3
\end{equation}
from three successive coefficients $b_{n-1}, \,\, b_n, \,\, b_{n+1}$, as before increasing the order of the lowest used coefficient by one until one runs out of coefficients. We do this below with $\mu$ varying within its estimated error range.
Fitting the available coefficients to the four unknowns, we estimate $c_1\approx 2.450 \pm 0.002,$ implying $\mu= 11.59 \pm 0.02,$ (in good agreement with our earlier estimate of 11.60), $c_2 =-3.23 \pm 0.03,$ implying $\mu_1 = 0.0396 \pm 0.0012,$ $c_3\approx -1,$ while $c_4$ is difficult to estimate beyond saying it is in the range $[1.3,3]$ implying $B \in [4,20].$
We repeated this analysis with a 3 parameter fit, varying $\mu$ in the range $[11.58,11.62].$ This gave $c_2 =-3.22 \pm 0.08,$ implying $\mu_1 = 0.040 \pm 0.003,$ $c_3=g= -1.15 \pm 0.2,$ and $c_4=1.8 \pm 0.5$ implying $B = 7 \pm 3.$
As noted above, differential approximants are useful insofar as they indicate that the singularity is not algebraic. They provide a signal, but are then of no further use in their current form. The presence of the $\mu_1^{\sqrt{n}}$ term is responsible for the lack of applicability of the method. However we can manipulate the series to remove the offending term, and then use this powerful method. From eqn. (\ref{logcan1}), defining $\tilde{b}_n = b_n/\sqrt{n},$ one has
\begin{equation} \label{eq:renorm}
c_n=2n^{3/2}(\tilde{b}_n - \tilde{b}_{n-1}) \sim (2g-\log{B}) + n\log(\mu)-g\log(n)
\end{equation}
So
$$d_n = \exp(c_n) \sim D\cdot \mu^n \cdot n^{-g} \cdot (1+ {\rm O}(n^{1-\sigma})),$$ with $D=\exp(2g)/B.$
The dominant asymptotics of the coefficients $d_n$ are now those of an algebraic singularity -- though the nature of the correction term, O$(n^{1-\sigma}),$ means that there is a confluent singularity with exponent less than 1. This means that the OGF $\sum d_n\cdot x^n$ can be analysed by standard methods used to analyse series with algebraic singularities. This includes the method of differential approximants and Bulirsch-Stoer extrapolation of ratios. For the former method, one should use 3rd order DAs, as the presence of a confluent singularity means that we require an ODE with two independent solutions, and to allow for a non-singular background term requires a third independent solution. For Bulirsch-Stoer extrapolation of ratios, the parameter $w=1$ should be used, as the ratios $d_n/d_{n-1} \sim \mu \cdot (1 -g/n + {\rm o}(1/n)).$
For the DA analysis we have used 3rd and 4th order ODEs. As an aside, we also tried 2nd order ODEs and these were unsatisfactory, as we expected, as they produced two singularities close together in an unsuccessful attempt at representing the confluent singularity. We summarise the results in table \ref{tab:ana}. The column labelled $L$ gives the degree of the inhomogeneous polynomial of the approximating ODEs. The entries give estimates, averaged over many approximants, of the position and exponent of the singularity of the ODEs. Full details of the method are given in \cite{G89, GJ09}. It is seen that the 3rd order DAs give estimates of the radius of convergence that are centred around $0.086140 = 1/11.609,$ with exponent estimates around $g \approx -0.93.$ The 4th order approximants give slightly higher estimates of both the critical point and the absolute value of the exponent. We estimate $
1/\mu \approx 0.08619,$ or $\mu \approx 11.602.$ This is remarkably close to the initial estimate above, $\mu =11.60.$ For the exponent $g$ we estimate $g = -1.0 \pm 0.1.$
\begin{table}
\caption{\label{tab:ana}
Critical point and exponent estimates for renormalised 1324 PAPs}
\begin{center}
\begin{tabular}{lllll} \hline \hline
$L$ & \multicolumn{2}{c}{Second order DA} &
\multicolumn{2}{c}{Third order DA} \\
\hline
& \multicolumn{1}{c}{$1/\mu$} & \multicolumn{1}{c}{$g-1$} &
\multicolumn{1}{c}{$1/\mu$} & \multicolumn{1}{c}{$g-1$} \\
\hline
0 & 0.086237& -2.160 & 0.086134& -1.925 \\
1 & 0.086116& -1.905 & 0.086156& -1.958 \\
2 & 0.086142& -1.942 & 0.086149& -1.944 \\
3 & 0.086159& -1.964 & 0.086167& -1.982 \\
4 & 0.086137& -1.926 & 0.086156& -1.960 \\
5 & 0.086111& -1.914 & 0.086162& -1.966 \\
6& 0.086110& -1.910 & 0.086158& -1.959 \\
7 & 0.086140& -1.926 & 0.086170& -1.980 \\
8 & 0.086143& -1.936& 0.086178& -1.997 \\
9 & 0.086145& -1.934 & 0.086186& -2.012 \\
10 & 0.086142& -1.931 & 0.086188& -2.011 \\
\hline \hline
\end{tabular}
\end{center}
\end{table}
We can also apply other standard techniques to the transformed series. The ratios of successive terms ($d_n$) of the transformed series when plotted against $1/n$ are now visually linear. Accordingly, we extrapolate the ratios of the coefficients of the transformed series using the Bulirsch-Stoer algorithm, with parameter $w=1.$ The results are shown in table \ref{tab:BS4}. The first two rows are behaving monotonically. The last entry in the second row suggests that $\mu < 11.622,$ and assessing the way entries in that row are decreasing, we judge the limit to be around $11.60 \pm 0.01.$ Together with the result of the DA analysis given above, we combine these two results and give as our final estimate $\mu=11.60 \pm 0.01.$ The quoted error is to be interpreted as a confidence interval, not a rigorous error bound. And, it should be stressed, our analysis is predicated on our assumption that $\sigma =1/2.$
\begin{table}
\caption{\label{tab:BS4}
Last seven entries in each row of the table of Bulirsch-Stoer extrapolants. Each successive row is the result of a successively higher degree of extrapolation. The ratios of successive coefficients of the transformed series are extrapolated here.}
\begin{center}
\begin{tabular}{llllllll} \hline \hline
L & T(L,N-L-6) & T(L,N-L-5) & T(L,N-L-4)& T(L,N-L-3) & T(L,N-L-2) & T(L,N-L-1) & T(L,N-L) \\
\hline
\hline
1&11.63526435& 11.63607983 &11.63648194& 11.63651932 &11.63624657& 11.63571995 &11.63499412\\
2&11.65200739 &11.64744495 &11.64238782 &11.63709746 &11.63181012 &11.62672996 &11.62202312 \\
3&11.64102224& 11.63870198& 11.63726970 &11.63654375& 11.63647001 &11.63709518& 11.63859791\\
\hline \hline
\end{tabular}
\end{center}
\end{table}
Finally, we estimate the amplitude $B$ by extrapolating the sequence $b_n/(\mu^n \cdot \mu_1^{\sqrt{n}} \cdot n^g),$ supplying the estimates of the critical parameters already found, against $1/n.$ In this way we estimate $B = 9.5 \pm 0.5.$
\section{Conclusion}
We have given a refined version of the JN algorithm that allows five further coefficients of the 1324 PAP generating function to be obtained, with comparable computer resources.
Analysing the coefficients of the generating function, we provide compelling evidence that they have a singularity structure of the form
$$B\cdot \mu^n \cdot \mu_1^{\sqrt{n}} \cdot n^g.$$ We give as our final estimates of the critical parameters $\mu=11.60 \pm 0.01,$ $\sigma=1/2,$ $\mu_1 = 0.0398 \pm 0.0010,$ $g = -1.1 \pm 0.2$ and $B =9.5 \pm 1.0.$ If, as is seen in other problems which have coefficients with a similar asymptotic structure, that $g$ is a simple rational fraction, the most likely is $-7/6$ or $-6/5,$ though we could not rule out $-1$ or even $-5/4.$
Zeilberger has said ``Not even God knows the number of 1324-avoiders of length 1000". While making no Messianic claims, our asymptotics permit the approximate answer $4.6 \times 10^{1017}.$
\section*{Acknowledgements}
AJG wishes to acknowledge helpful conversations with Einar Steingr\'imsson, who brought this problem to our attention, and the hospitality of Mathematisches Forschungsinstitut Oberwolfach and the Enumerative Combinatorics Workshop held there on March 2--8, 2014, where this work was initiated. We are grateful to Alan Sokal, who gave us access to his Dell computers with 1TB memory, which were needed for these calculations, through his NSF grant PHY--0424082 NYU. AJG wishes to thank the Australian Research Council for supporting this work through grant DP120100931.
|
\section{Introduction}
In 2010, Winter and Zacharias introduced the notion of nuclear dimension for \linebreak $C^*$-algebras as a non-commutative analogue of topological covering dimension \linebreak (\cite{WZ10}).
Since then it has clearly become one of the most important concepts \linebreak in the Elliott classification program, i.e. the classification of simple, nuclear \linebreak $C^*$-algebras by $K$-theoretic data.
In particular, finite-dimensionality with respect to this notion of dimension constitutes one of the fundamental regularity properties for the $C^*$-algebras in question.
It is part of the Toms-Winter conjecture that in this situation finite nuclear dimension is presumably equivalent to $\mathcal{Z}$-stability and therefore a necessary assumption in order to obtain classification by the Elliott invariant.
This conjecture received a great deal of attention and the equivalence mentioned above has been partially verified (\cite{SWW14}, \cite{Win12}).
On the other hand, large classes of simple, nuclear $C^*$-algebras with finite nuclear dimension have been successfully classified by their Elliott invariant (e.g. \cite{Lin11}, \cite{Win12}).
However, once a $C^*$-algebra having finite nuclear dimension belongs to a class which is classified by some invariant, it is a natural question to ask for the precise value of its dimension and how it can be read off from its invariant.
In this paper we study this question for the class of UCT-Kirchberg algebras, i.e. purely infinite, simple, separable, nuclear $C^*$-algebras satisfying the UCT which are, due to Kirchberg and Phillips, completely classified by their $K$-groups.
In \cite{WZ10}, Winter and Zacharias showed that all these algebras have nuclear dimension at most 5 and asked whether the precise value of their dimension is determined by algebraic properties of their $K$-groups, such as torsion (\cite[Problem 9.2]{WZ10}).
Here, we give a partial answer to their question and give the optimal dimension estimate in the absence of $K_1$-torsion.
More precisely, we show that the nuclear dimension of UCT-Kirchberg algebras with torsion free $K_1$-groups equals 1 (Theorem \ref{main}).
Our methods of proof used in the computation of the nuclear dimension differ from the arguments in the original estimate by Winter and Zacharias in two essential ways.
First, we do not make use of Cuntz-Pimsner algebras, as done in \cite{WZ10}, but employ crossed product models for UCT-Kirchberg algebras which are provided by the work of R{\o}rdam (\cite{Ror95}).
Second, we do not construct completely positive approximations for the identity map on a given Kirchberg algebra $A$ directly.
Instead, we study a certain family of homomorphisms $\iota_n\colon A\rightarrow M_n(A)$ and show that the $\iota_n$ can be approximated in a decomposable manner.
While the error we have to make in the approximation of each $\iota_n$ is bounded from below, the lower bound will be small for large $n$.
In a second step, we show how to combine the $\iota_n$ in order to construct a homomorphism $A\rightarrow M_k(A)$ which admits decomposable approximations of the same quality as the $\iota_n$, but in addition induces an isomorphism on $K$-theory.
By Kirchberg-Phillips classification, this map can be perturbed to the identity map on $A$ and by that provides suitable approximations for estimating the nuclear dimension of $A$. \vspace{1cm}
This paper is organized as follows: First, right now, we define certain matrix embeddings for crossed products which are the main objects of study in this paper.
\begin{definition}\label{iota}
Let $A$ be a $C^*$-algebra, $\alpha\in\Aut(A)$ an automorphism and $A\rtimes_\alpha\mathbb{Z}$ the corresponding full crossed product.
For $n\in\mathbb{N}$, $n\geq 2$, we let
\[\iota_n\colon A\rtimes_\alpha\mathbb{Z} \rightarrow M_n(A\rtimes_\alpha\mathbb{Z})\]
denote the embedding given by
\[\begin{tabular}{ccc}$a \mapsto\begin{pmatrix}
\alpha^{-1}(a) & 0 & \cdots& 0 \\ 0 & \alpha^{-2}(a) & \ddots& \vdots\\ \vdots & \ddots& \ddots& 0 \\ 0 & \cdots& 0 & \alpha^{-n}(a)
\end{pmatrix}$
& and
& $U \mapsto\begin{pmatrix}
0 & \cdots & \cdots & 0 & U^n \\
1 & \ddots &&& 0 \\
0 & \ddots & \ddots & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0 & \cdots & 0 & 1 &0
\end{pmatrix}$\end{tabular}\]
for $a\in A$ resp. for the unitary $U\in\mathcal{M}(A\rtimes_\alpha\mathbb{Z})$ implementing the action $\alpha$.
\end{definition}
In section \ref{section approximation}, we show that these maps $\iota_n$ admit completely positive approximations in an order zero fashion, at least up to an error which is small for large $n$.
The $K$-theory of these maps can be computed in many cases of interest, this is done in section \ref{section k-theory}.
A combination of these results yields the dimension estimate for Kirchberg algebras in section \ref{section main}.
We finish with some remarks about possible generalizations of the technique developed in this paper.
\section{Decomposable approximations for $\iota_n$}\label{section approximation}
We want to study certain maps which admit completely positive approximations of the same type as in the original definition of nuclear dimension in \cite{WZ10}.
In our situation, however, we won't be able to find approximations of arbitrary small error (which would rather lead to a notion of nuclear dimension for maps).
Therefore we need to keep track of the precision of approximation that we can get for a given map.
This will be done using the following definition.
\begin{definition}\label{def decomposition}
Let $\alpha\colon A\rightarrow B$ be a map between $C^*$-algebras $A$ and $B$.
Given \linebreak $d\in\mathbb{N}_0$, $\epsilon>0$ and a finite subset $\mathcal{G}$ of $A$, we say that $\alpha$ admits a piecewise contractive, completely positive, $d$-decomposable $(\mathcal{G},\epsilon)$-approximation if
there exists $(F,\psi,\varphi)$ such that $F$ is a finite-dimensional $C^*$-algebra, and such that $\psi\colon A\rightarrow F$ and $\varphi\colon F\rightarrow B$ are completely positive maps satisfying
\begin{enumerate}
\item $\|(\varphi\circ\psi)(a)-\alpha(a)\|<\epsilon$ for all $a\in\mathcal{G}$;
\item $\psi$ is contractive;
\item $F$ decomposes as $F=F^{(0)}\oplus\cdots\oplus F^{(d)}$ such that $\varphi_{|F^{(i)}}$ is a c.p.c. order zero map for each $i=0,...,d$.
\end{enumerate}
\end{definition}
In this terminology, the notion of nuclear dimension for a $C^*$-algebra $A$, as defined by Winter and Zacharias in \cite{WZ10}, can be formulated as follows:
\[\begin{tabular}{rc}
\multirow{2}{*} {$\nucdim(A)\leq d\quad\Leftrightarrow$} &
$\id_A$ admits piecewise contractive, c.p., $d$-decomposable, \\ & $(\mathcal{G},\epsilon)$-approximations for all finite $\mathcal{G}\subset A$ and all $\epsilon>0$.
\end{tabular}\]
We first note that the approximations of \ref{def decomposition} for $^*$-homomorphisms are well-behaved with respect to composition, orthogonal sum and approximate unitary equivalence.
We collect some of these elementary permanence results in the following lemma.
\begin{lemma}\label{permanence}
Let $*$-homomorphisms $\xymatrix{A \ar[r]^\alpha & B \ar[r]^{\beta_1,\beta_2} & C\ar[r]^\gamma & D}$ between \linebreak $C^*$-algebras $A,B,C,D$, a finite subset $\mathcal{G}\subset B$, $\epsilon>0$ and $d\in\mathbb{N}_0$ be given.
If both $\beta_1$ and $\beta_2$ admit piecewise contractive, completely positive, $d$-decomposable $(\mathcal{G},\epsilon)$-approximations $(F_i,\psi_i,\varphi_i)$, $i\in\{1,2\}$, then the following holds:
\begin{enumerate}
\item The composition $\gamma\circ\beta_1$ admits a piecewise contractive, completely positive, $d$-decomposable $(\mathcal{G},\epsilon)$-approximation.
\item For any $\alpha$-preimage $\mathcal{G}'$ of $\mathcal{G}$, the composition $\beta_1\circ\alpha$ admits a piecewise contractive, completely positive, $d$-decomposable $(\mathcal{G'},\epsilon)$-approximation.
\item The orthogonal sum $\begin{pmatrix}\beta_1 & 0 \\ 0 & \beta_2\end{pmatrix}\colon B\rightarrow M_2(C)$ admits a piecewise contractive, completely positive, $d$-decomposable $(\mathcal{G},\epsilon)$-approximation.
\item If $\beta_1\sim_{a.u.}\delta$ for some $\delta\colon B\rightarrow C$, then $\delta$ admits a piecewise contractive, completely positive, $d$-decomposable $(\mathcal{G},\epsilon)$-approximation.
\end{enumerate}
\end{lemma}
\begin{proof}
We only name suitable approximating systems and leave the details to the reader.
For (1) one chooses $(F_1,\psi_1,\gamma\circ\varphi_1)$, while for (2) $(F_1,\psi_1\circ\alpha,\varphi_1)$ works.
For (3) consider $\left(F_1\oplus F_2,\psi_1\oplus\psi_2,\begin{pmatrix}\varphi_1 & 0 \\ 0 & \varphi_2\end{pmatrix}\right)$ with respect to the decomposition $(F_1\oplus F_2)^{(i)}=F_1^{(i)}\oplus F_2^{(i)}$.
Given $\delta=\lim\limits_{n\rightarrow\infty}\Ad(U_n)\circ\beta_1$ in (4), one checks that the triple $(F,\psi_1,\Ad(U_n)\circ\varphi_1)$ works for $n$ sufficiently large.
\end{proof}
Next, we apply a 'cut and paste'-technique similar to the one used in \cite[section 7]{WZ10} to the maps $\iota_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ of Definition \ref{iota}.
This essentially reduces a decomposable approximation of $\iota_n$ to a decomposable approximation of the coefficient algebra $A$.
\begin{lemma}\label{approximation1}
Let a nuclear $C^*$-algebra $A$, an automorphism $\alpha\in\Aut(A)$, a finite subset $\mathcal{G}\subset A\rtimes_\alpha\mathbb{Z}$ and $\epsilon>0$ be given.
Then the following holds for the maps $\iota_n$ as defined in \ref{iota}:
For all sufficiently large $n$ there exists a c.c.p. map $\psi_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A)$ and two $\ast$-homomorphisms $\Lambda_n^0,\Lambda_n^1\colon M_n(A)\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ such that the diagram
\[
\begin{xy}
\xymatrix{A\rtimes_\alpha\mathbb{Z} \ar[rr]^{\iota_n} \ar[dr]_{\psi_n} && M_n(A\rtimes_\alpha\mathbb{Z}) \\ & M_n(A) \ar[ur]_{\Lambda_n^0+\Lambda_n^1}}
\end{xy}
\]
commutes up to $\epsilon$ on $\mathcal{G}$, i.e. $\|\iota_n(x)-((\Lambda_n^0+\Lambda_n^1)\circ\psi_n)(x)\|<\epsilon$ holds for all $x\in\mathcal{G}$.
\end{lemma}
\begin{proof}
Assume $A\subseteq\mathcal{B}(H)$ for some Hilbert space $H$.
By nuclearity of $A$ we have $A\rtimes_\alpha\mathbb{Z}=A\rtimes_{\alpha,r}\mathbb{Z}$ (\cite[Theorem 4.2.4]{BO08}) and may therefore identify the crossed product with the $C^*$-subalgebra of $\mathcal{B}(\ell^2(\mathbb{Z})\otimes H)$ generated by the operators
\[\begin{array}{rclr}
a(\delta_i\otimes\xi)&=&\delta_i\otimes(\alpha^{-i}(a))(\xi),&a\in A \\
U(\delta_i\otimes\xi)&=&\delta_{i+1}\otimes\xi
\end{array}\]
respectively by the ideal therein generated by $A$ in the non-unital case.
Denote by $P_n$ the projection onto the subspace $\ell^2(\{1,...,n\})\otimes H$.
Compression by $P_n$ gives a c.p.c. map from $A\rtimes_\alpha\mathbb{Z}$ to $M_n(A)$.
We need to put suitable weights on the entries of the elements of $\{P_nxP_n\}_{x\in\mathcal{G}}$.
As in \cite{WZ10}, this will be done by Schur multiplication (i.e. entrywise multiplication) with a suitable positive contraction $\kappa_n\in M_n(\mathbb{C})$.
We choose
\[\kappa_n=\frac{1}{m+1}\begin{pmatrix}
1 & 1 & \cdots & 1 & 1 & 1 & 1 & \cdots & 1 & 1 \\
1 & 2 & \cdots & 2 & 2 & 2 & 2 & \cdots & 2 & 1 \\
\vdots & \vdots & \ddots & \vdots & \vdots & \vdots & \vdots & \iddots & \vdots & \vdots \\
1 & 2 & \cdots & m-1 & m-1 & m-1 & m-1 & \cdots & 2 & 1 \\
1 & 2 & \cdots & m-1 & m & m & m-1 & \cdots & 2 & 1 \\
1 & 2 & \cdots & m-1 & m & m & m-1 & \cdots & 2 & 1 \\
1 & 2 & \cdots & m-1 & m-1 & m-1 & m-1 & \cdots & 2 & 1 \\
\vdots & \vdots & \iddots & \vdots & \vdots & \vdots & \vdots & \ddots & \vdots & \vdots \\
1 & 2 & \cdots & 2 & 2 & 2 & 2 & \cdots & 2 & 1 \\
1 & 1 & \cdots & 1 & 1 & 1 & 1 & \cdots & 1 & 1
\end{pmatrix}\]
if $n=2m$ is even, respectively
\[\kappa_n=\frac{1}{m+1}\begin{pmatrix}
1 & 1 & \cdots & 1 & 1 & 1 & \cdots & 1 & 1 \\
1 & 2 & \cdots & 2 & 2 & 2 & \cdots & 2 & 1 \\
\vdots & \vdots & \ddots & \vdots & \vdots & \vdots & \iddots & \vdots & \vdots \\
1 & 2 & \cdots & m & m & m & \cdots & 2 & 1 \\
1 & 2 & \cdots & m & m+1 & m & \cdots & 2 & 1 \\
1 & 2 & \cdots & m & m & m & \cdots & 2 & 1 \\
\vdots & \vdots & \iddots & \vdots & \vdots & \vdots & \ddots & \vdots & \vdots \\
1 & 2 & \cdots & 2 & 2 & 2 & \cdots & 2 & 1 \\
1 & 1 & \cdots & 1 & 1 & 1 & \cdots & 1 & 1
\end{pmatrix}\]
if $n=2m+1$ is odd.
It is not hard to see that the map on $M_n(A)$ given by $y\mapsto\kappa_n\ast y$, where $\ast$ denotes Schur multiplication, is then in fact completely positive and contractive.
Therefore, $\psi_n$ defined as
\[\begin{array}{cc}\psi_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A),&x\mapsto\kappa_n\ast(P_nxP_n)\end{array}\]
is also a c.p.c. map.
Now let $j\colon M_n(A)\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ be the canonical embedding. Using the unitary element
\[V:=\begin{pmatrix}
0 & \cdots & \cdots & 0 & U^{n-1} \\
U^{-1} & \ddots &&& 0 \\
0 & \ddots & \ddots & & \vdots \\
\vdots & \ddots & \ddots & \ddots & \vdots \\
0 & \cdots & 0 & U^{-1} &0
\end{pmatrix}\in M_n(\mathcal{M}(A\rtimes_\alpha\mathbb{Z})),\]
we consider the $^*$-homomorphisms $\Lambda_n^0,\Lambda_n^1\colon M_n(A)\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ given by $\Lambda_n^0=j$ and $\Lambda_n^1=Ad\left(V^m\right)\circ j$.\\
We may assume that the elements of $\mathcal{G}$ are of the form $x=a_xU^{k_x}$ for suitable contractions $a_x\in A$ and $k_x\in\mathbb{Z}$, let $K:=\max_{x\in\mathcal{G}}|k_x|$.
We claim that, for $n>\frac{K}{\epsilon}$, the maps $\Lambda_n^0,\Lambda_n^1$ and $\psi_n$ defined above provide an approximation of $\iota_n$ as claimed.
For convenience we only consider the case of even $n$, the calculations for the odd case are essentially identical.
Given $x=aU^{k}\in\mathcal{G}$ with $k\geq 0$, one checks that $\iota_n(x)-((\Lambda_n^0+\Lambda_n^1)\circ\psi_n)(x)$ equals
\[\frac{1}{(\frac{n}{2}+1)}\begin{pmatrix}
0&0&0&k\cdot I_k-\mu_k \\ k\cdot I_{(\frac{n}{2}-k)} &0&0&0\\ 0& k\cdot I_k - \mu_k &0&0 \\0&0&k\cdot I_{(\frac{n}{2}-k)}&0
\end{pmatrix}\ast\iota_n(x)\]
where
\[\mu_k=\begin{pmatrix}
0 & 0 & \cdots &&& \cdots & 0 \\
0 & 1 & 0 &&&& \vdots \\
\vdots & 0 & 2 & \ddots \\
&&\ddots & \ddots & \ddots \\
&&&\ddots & 2 & 0 & \vdots \\
\vdots &&&&0&1&0\\
0&\cdots&&&\cdots&0&0\end{pmatrix}\in M_k(\mathbb{C}),\]
$I_d$ is the $d\times d$ unit matrix and $\ast$ denotes again Schur multiplication.
Due to the special off-diagonal form of these elements, an entrywise norm estimate immediately shows
\[\|\iota_n(x)-((\Lambda_n^0+\Lambda_n^1)\circ\psi_n)(x)\|\leq\frac{k}{(\frac{n}{2}+1)}\|x\|\leq\frac{K}{n}<\epsilon.\]
\end{proof}
Next, we show that the maps $\iota_n$ of \ref{iota} admit decomposable approximations in the sense of Definition \ref{def decomposition} provided that the coefficient algebra has finite nuclear dimension.
\begin{proposition}\label{approximation2}
Given a $C^*$-algebra $A$ with $\nucdim(A)=d<\infty$, a finite subset $\mathcal{G}\subset A$ and $\epsilon>0$, the maps $\iota_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ defined in \ref{iota} admit a piecewise contractive, completely positive, $(2d+1)$-decomposable $(\mathcal{G},\epsilon)$-approximation for all sufficiently large $n$.
\end{proposition}
\begin{proof}
Choose $n$ large enough such that there exist maps $\psi_n$ and $\Lambda_n^0,\Lambda_n^1$ as in Lemma \ref{approximation1} satisfying $\|\iota_n(x)-\left((\Lambda_n^0+\Lambda_n^1)\circ\psi_n\right)(x)\|<\frac{\epsilon}{3}$ for all $x\in\mathcal{G}$.
Since $\nucdim(M_n(A))=\nucdim(A)=d$ by \cite[Corollary 2.8]{WZ10}, there exists a piecewise contractive, completely positive, $d$-decomposable $(\psi_n(\mathcal{G}),\frac{\epsilon}{3})$-approximation of $\id_{M_n(A)}$,
i.e. a finite-dimensional $C^*$-algebra $F=F^{(0)}\oplus...\oplus F^{(d)}$, a c.p.c. map $\psi'\colon M_n(A)\rightarrow F$ and c.p.c. order zero maps $\varphi'^{(i)}\colon F^{(i)}\rightarrow M_n(A)$, $i=0,...,d$, such that
\[\left\|\left(\sum_{i=0}^d\varphi'^{(i)}\circ\psi'\right)(\psi_n(x))-\psi_n(x)\right\|<\frac{\epsilon}{3}\] holds for all $x\in\mathcal{G}$.
Putting these two approximations together, i.e. setting $G:=F\oplus F$ with decomposition $G=\bigoplus_{j=0,1}\bigoplus_{i=0}^d G^{(i,j)}$, where $G^{(i,j)}:=F^{(i)}$,
and considering
\[\begin{xy}\xymatrix{
A\rtimes_\alpha\mathbb{Z} \ar[rr]^{\iota_n} \ar[dr]_(0.4){(\psi'\circ\psi_n)\oplus(\psi'\circ\psi_n)\;\;\;\;\;} && M_n(A\rtimes_\alpha\mathbb{Z}) \\
& G\oplus G=\bigoplus\limits_{j=0,1}\;\bigoplus\limits_{i=0}^d G^{(i,j)} \ar[ur]_(0.7){\sum\limits_{j=0,1}\;\sum\limits_{i=0}^d\left(\Lambda_n^j\circ\varphi'^{(i)}\right)}
}\end{xy},\]
one finds each $\Lambda_n^j\circ\varphi'^{(i)}$ to be an order zero map and further
\[\begin{array}{rl}&\left\|\sum\limits_{j=0}^1\sum\limits_{i=0}^d\left(\Lambda_n^j\circ\varphi'^{(i)}\right)((\psi'\circ\psi_n)(x))-\iota_n(x)\right\| \\ \\
\leq & \left\|\sum\limits_{j=0}^1\Lambda_n^j(\psi_n(x))-\iota_n(x)\right\| +2\left\|\left(\sum\limits_{i=0}^d\varphi'^{(i)}\circ\psi'\right)(\psi_n(x))-\psi_n(x)\right\| \\ \\ < & \epsilon
\end{array}\]
for all $x\in\mathcal{G}$.
In other words, $\left(G,(\psi'\circ\psi_n)^{\oplus 2},\sum_{i,j}\left(\Lambda_n^j\circ\varphi'^{(i)}\right)\right)$ provides a piecewise contractive, completely positive, $(2d+1)$-decomposable $(\mathcal{G},\epsilon)$-approximation for $\iota_n$.
\end{proof}
\section{The $K$-theory of $\iota_n$}\label{section k-theory}
In this section we compute the map on $K$-theory induced by $\iota_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ from \ref{iota}.
Using a rotation argument we first express $K_*(\iota_n)$ in terms of $K_*(\alpha)$.
Under suitable assumptions, this allows us to read off $K_*(\iota_n)$ from the Pimsner-Voiculescu sequence associated to $(A,\alpha)$.
\begin{lemma}\label{homotopy}
The homomorphism $\iota_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ defined in \ref{iota} is homotopic to the diagonal embedding $j_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ given by
\[\begin{tabular}{ccc}$a \mapsto\begin{pmatrix}
\alpha^{-1}(a) & 0 & \cdots& 0 \\ 0 & \alpha^{-2}(a) & \ddots& \vdots\\ \vdots & \ddots& \ddots& 0 \\ 0 & \cdots& 0 & \alpha^{-n}(a)
\end{pmatrix}$
& and
& $U \mapsto\begin{pmatrix}
U \\ & U \\ &&\ddots \\&&&U
\end{pmatrix}$\end{tabular}\]
for $a\in A$ resp. for the unitary $U\in\mathcal{M}(A\rtimes_\alpha\mathbb{Z})$ implementing the action $\alpha$.
\end{lemma}
\begin{proof}
Let $(V_t)_{t\in[0,1]}$ be a path of unitaries in $M_n(\mathbb{C})$ connecting the shift-unitary $V_0=\sum_{i=1}^{n-1} e_{i,i+1}+e_{n,1}$ to identity matrix $V_1=1_n$ and consider the path $(\kappa_t)_{t\in[0,1]}$ of homomorphisms $\kappa_t\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ given by
\[\begin{tabular}{ccc}$a \mapsto\begin{pmatrix}
\alpha^{-1}(a) & 0 & \cdots& 0 \\ 0 & \alpha^{-2}(a) & \ddots& \vdots\\ \vdots & \ddots& \ddots& 0 \\ 0 & \cdots& 0 & \alpha^{-n}(a)
\end{pmatrix}$
& and
& $U \mapsto V_t\ast\begin{pmatrix}
U & U^2 & \cdots & U^n \\
1 & U & \ddots & \vdots \\
\vdots & \ddots & \ddots & U^2 \\
U^{-n+2} & \cdots & 1 & U
\end{pmatrix}$\end{tabular}\]
where $\ast$ denotes Schur multiplication.
It is straightforward to check that $\kappa_t(U)$ is a continuous unitary path in $M_n(\mathcal{M}(A\rtimes_\alpha\mathbb{Z}))$, hence we only have to make sure that each $\kappa_t$ is well-defined, i.e. compatible with the action $\alpha$:
\[\begin{array}{rl}
&\left(\kappa_t(U)\kappa_t(a)\kappa_t(U)^*\right)_{ij} \\
= & \sum_{k=1}^n \kappa_t(U)_{ik}\alpha^{-k}(a)\kappa_t(U^*)_{kj} \\
= & \sum_{k=1}^n (V_t)_{ik}\overline{(V_t)_{jk}}U^{(k-i+1)}\alpha^{-k}(a)U^{-(k-j+1)} \\
= & \left(\sum_{k=1}^n (V_t)_{ik}\overline{(V_t)_{jk}}\right) U^{-i+1}aU^{j-1} \\
= & \delta_{ij}\cdot U^{-i+1}aU^{j-1} \\
= & \delta_{ij}\cdot \alpha^{-i+1}(a) \\
= & \left(\kappa_t(\alpha(a))\right)_{ij}
\end{array}\]
This shows that $\kappa_t$ provides a homotopy between $\kappa_0=\iota_n$ and $\kappa_1=j_n$.
\end{proof}
\begin{proposition}\label{k-theory}
Given $n\in\mathbb{N}$, let $\iota_n\colon A\rtimes_\alpha\mathbb{Z}\rightarrow M_n(A\rtimes_\alpha\mathbb{Z})$ be the homomorphism defined in \ref{iota}.
If both boundary maps in the Pimsner-Voiculescu sequence associated to $(A,\alpha)$ vanish, we find $K_*(\iota_n)=n\cdot \id_{K_*(A\rtimes_\alpha\mathbb{Z})}$.
The same conclusion holds if instead either $K_0(A)=0$ or $K_1(A)=0$.
\end{proposition}
\begin{proof}
We know by Lemma \ref{homotopy} that $(\iota_n)_*=\sum_{i=1}^n (\widehat{\alpha^i})_*$ where $\hat{\beta}$ denotes the canonical extension of an automorphism $\beta\in\Aut(A)$ satisfying $\alpha\circ\beta=\beta\circ\alpha$ to an automorphism of $A\rtimes_\alpha\mathbb{Z}$ via $\widehat{\beta}(U)=U$.
Using $(\widehat{\alpha^i})_*=(\widehat{\alpha}_*)^i$, it suffices to show that $\widehat{\alpha}$ induces the identity map on $K$-theory.
By the Pimsner-Voiculescu sequence (\cite[Theorem 10.2.1]{Bla98}) and the naturality thereof, we have the following commutative diagram with exact rows
\[\begin{xy}\xymatrix{
\ar[r] & K_*(A) \ar[r]^{1-\alpha_*} \ar[d]^{\alpha_*} & K_*(A) \ar[r]^(.4){i_*} \ar[d]^{\alpha_*} & K_*(A\rtimes_\alpha\mathbb{Z}) \ar[r]^{\partial_*} \ar[d]^{\widehat{\alpha}_*} & K_{*+1}(A) \ar[r]^{1-\alpha_{*+1}} \ar[d]^{\alpha_{*+1}} & K_{*+1}(A) \ar[r] \ar[d]^{\alpha_{*+1}} & \\
\ar[r] & K_*(A) \ar[r]^{1-\alpha_*} & K_*(A) \ar[r]^(.4){i_*} & K_*(A\rtimes_\alpha\mathbb{Z}) \ar[r]^{\partial_*} & K_{*+1}(A) \ar[r]^{1-\alpha_{*+1}} & K_{*+1}(A) \ar[r] & }\end{xy}\]
where $i$ denotes the canonical inclusion of $A$ into the crossed product.
In the first case, i.e. $\partial_*=0$, we have an isomorphism $K_*(A)/(1-\alpha_*)K_*(A)\cong K_*(A\rtimes_\alpha\mathbb{Z})$ induced by $i_*$.
Hence $\widehat{\alpha}_*\circ i_*=i_*\circ\alpha_*=i_*$ and the claim follows by surjectivity of $i_*$.
The second case is treated similarly. If $K_i(A)=0$, then $K_{i+1}(A\rtimes_\alpha\mathbb{Z})\cong K_{i+1}(A)/(1-\alpha_{i+1})K_{i+1}(A)$ and $\widehat{\alpha}_{i+1}=\id$ follows as in the first case.
On the other hand, the boundary map $\partial_i$ identifies $K_i(A\rtimes_\alpha\mathbb{Z})$ with $ker(1-\alpha_{i+1})\subseteq K_{i+1}(A)$.
Hence $\partial_i\circ\widehat{\alpha}_i=\alpha_{i+1}\circ\partial_i=\partial_i$, which shows that also $\widehat{\alpha}_i=\id$.
\end{proof}
\section{Nuclear dimension of Kirchberg algebras}\label{section main}
Combining the results of section \ref{section approximation} and \ref{section k-theory} with the $K$-theoretical classification results by Kirchberg and Phillips, we obtain our main result, the computation of the nuclear dimension for UCT-Kirchberg algebras in the absence of torsion in $K_1$.
\begin{theorem}\label{main}
Let $B$ be a Kirchberg algebra in the UCT-class. If $K_1(B)$ is torsion free, then $\nucdim(B)=1$.
\end{theorem}
\begin{proof}
Since $\nucdim(B)=\nucdim(B\otimes\mathbb{K})$ by \cite[Corollary 2.8]{WZ10}, we may assume that $B$ is stable.
In this case, $B$ can be realized as a crossed product $B=A\rtimes_\alpha\mathbb{Z}$ with $A$ an AF-algebra by \cite[Corollary 8.4.11]{Ror02}.
This implies that, given any finite subset $\mathcal{G}$ of $B=A\rtimes_\alpha\mathbb{Z}$ and $\epsilon>0$, we can use Proposition \ref{approximation2} to find a natural number $n$ such that both $\iota_n\colon B\rightarrow M_n(B)$ and $\iota_{n+1}\colon B\rightarrow M_{n+1}(B)$ (as defined in \ref{iota}) admit piecewise contractive, completely positive, 1-decomposable $(\mathcal{G},\epsilon)$-approximations $(F_i,\psi_i,\varphi_i)$, $i\in\{n,n+1\}$.
Using the existence part of Kirchberg-Phillips classification (\cite[Theorem 8.4.1]{Ror02}), we can choose an automorphism $\omega$ on $M_n(B)$ which induces $\omega_*=-\id$ on $K$-theory.
Now consider the embedding $\iota$ given by
\[\iota=\begin{pmatrix}\iota_{n+1} & 0 \\ 0 & \omega\circ\iota_n\end{pmatrix}\colon B\rightarrow M_{2n+1}(B).\]
By Lemma \ref{permanence}, $\iota$ also admits a piecewise contractive, completely positive, 1-decom\-posable $(\mathcal{G},\epsilon)$-approximation.
By Proposition \ref{k-theory} and the choice of $\omega$ we further have
\[K_*(\iota)=K_*(\iota_{n+1})+K_*(\omega)\circ K_*(\iota_n)=(n+1)\cdot\id-n\cdot\id=\id.\]
Since $B$ satisfies the UCT, this shows that the class of $\iota$ is invertible in \linebreak $KK(B,M_{2n+1}(B))$ and its inverse is, using \cite[Theorem 8.4.1]{Ror02} again, induced by a $*$-isomomorphism $\varrho\colon M_{2n+1}(B)\rightarrow B$.
The uniqueness part of Kirchberg-Phillips classification (\cite[Theorem 8.4.1]{Ror02}) implies that $\varrho\circ\iota$ is approximately unitarily equivalent to $\id_B$.
Hence, by Lemma \ref{permanence}, $\id_B$ also admits a piecewise contractive, completely positive, 1-decomposable $(\mathcal{G},\epsilon)$-approximation.
Since $\mathcal{G}$ and $\epsilon$ were arbitrary, this shows $\nucdim(B)\leq 1$.
Because $B$ is not an AF-algebra, the nuclear dimension of $B$ must in fact be equal to 1.
\end{proof}
For the general case, i.e. for Kirchberg algebras without any $K$-theory constraints, we immediately get the following.
Note that the same estimate has by now also be obtained in \cite[Theorem 7.1]{MS13} and \cite[Corollary 3.4]{BEMSW14} by methods completely different from the one developed here.
These results are more general though, they do not require the UCT.
\begin{corollary}
A Kirchberg algebra in the UCT-class has nuclear dimension at most 3.
\end{corollary}
\begin{proof}
This is proven exactly as in \cite[Theorem 7.5]{WZ10} except for using the improved estimate $\nucdim(\mathcal{O}_\infty)=1$ which is given by Theorem \ref{main}.
\end{proof}
The strategy of this paper can be used to obtain dimension estimates in more general situations (cf. Remark \ref{Efren} below).
In fact, there are only two ingredients needed:
First, one needs sufficient classfication results, i.e. both existence and uniqueness results for homomorphisms between $C^*$-algebras in the class under consideration.
Second, one requires the existence of a family of homomorphisms which admit good approximations in a decomposable manner.
These homomorphisms should further induce maps on the level of invariants which can be used to build up an invertible element.
In that case, one can proceed as in Theorem \ref{main} to obtain dimension estimates.
However, the proof of \ref{main} seems to be limited to infinite $C^*$-algebras.
This is because we require an automorphism $\omega$ which gives a sign on $K$-theory, i.e. which satisfies $\omega_*=-\id$.
However, the map $-\id$ will not be an automorphism as soon as the invariant involves some non-trivial order structure on the $K$-groups.
Therefore one cannot find such $\omega$ in this case.
\begin{remark}\label{Efren}
After having recieved a preprint of this paper, Ruiz, Sims and Tomforde estimated the nuclear dimension of certain graph $C^*$-algebras in \cite{RST13}, following the strategy outlined above.
Their result includes some, but not all Kirchberg algebras covered by Theorem \ref{main}.
Moreover, using classification results involving filtered-$K$-theory they also obtain estimates in the non-simple case.
\end{remark}
\begin{remark}\label{AT-case}
In the presence of torsion in $K_1$, one can still represent stable UCT-Kirchberg algebras $B$ as crossed products $A\rtimes_\alpha\mathbb{Z}$.
We can no longer choose $A$ to be an AF-algebra, but one can settle for $A$ to be an A$\mathbb{T}$-algebra instead.
Furthermore, the crossed product can be constructed in such a way that the boundary maps in the associated Pimsner-Voiculescu sequence vanish (Theorem 3.6 of \cite{Ror95} and its proof) and therefore Proposition \ref{k-theory} applies.
Following the lines of \ref{main} and using nuclear one-dimensionality of A$\mathbb{T}$-algebras, this gives an alternative proof for the general estimate $\nucdim(B)\leq 3$.
However, in this case it looks like the maps $\iota_n$ from \ref{iota} even admit $2$-decomposable approximations when contructed more carefully.
This would give an improved dimension estimate in the torsion case.
\end{remark}
|
\section{Introduction}
The Discrete Markus-Yamabe Question is a problem concerning discrete dynamics, formulated in dimension $n$ by
Cima {\em et al.} \cite{Cima-Manosa} as follows:
\medbreak
\noindent{\em
{\bf [DMYQ($n$)]}
Let $f:\mathbf{R}^n\longrightarrow\mathbf{R}^n$ be a $C^1$ map such that $f(0)=0$ and for any $x\in\mathbf{R}^n$, $ Jf(x)$ has all its eigenvalues with modulus less than one.
Is it true that $0$ is a global attractor for the discrete dynamical system generated by $f$?
}
\medbreak
It is known that the answer is affirmative in dimension $1$ and there are counter-examples for dimensions higher than 2, see Cima {\em et al.} \cite{CvdEGHM} and van den Essen and Hubbers \cite{vdEH}.
In dimension 2, Cima {\em et al.} \cite{Cima-Manosa} prove that an affirmative answer is obtained when $f$ is a polynomial map, and provide a counter example which is a rational map.
After this, research on planar maps focused on the quest for minimal sufficient conditions under which the DMYQ has an affirmative answer.
Alarc\'on {\em et al.} \cite{Alarcon} use the existence of an invariant embedded curve joining the origin to infinity to show the global stability of the origin.
Symmetry is a natural context for the existence of such a curve, and this led us to a symmetric approach to this problem and to the results in \cite{AlarconDenjoy,SofisbeSzlenk,SofisbePolynomial,SofisbeGlobal,SofisbeSaddle} that we review in this article.
The present article studies maps $f$ of the plane which preserve symmetries described by the action of a compact Lie group.
In this setting we characterise the possible local dynamics near the unique fixed point of $f$, that we assume hyperbolic.
We establish for which symmetry groups local dynamics extends globally.
For the remaining groups we present illustrative examples.
\section{Preliminaries} \label{secPre}
This section consists of definitions and known results about topological dynamics and equivariant theory. These are grouped in two separate subsections, which are elementary for readers in each field,
containing material from the corresponding sections of \cite{AlarconDenjoy,SofisbeSzlenk,SofisbePolynomial,SofisbeGlobal,SofisbeSaddle} and is included here for ease of reference.
\subsection{Topological Dynamics}
We consider planar topological embeddings, that is, continuous and injective maps defined in $\mathbf{R}^ 2$. The set of topological embeddings of the plane is denoted by $\mbox{Emb}(\mathbf{R}^2)$.
Recall that for $f\in \mbox{Emb}(\mathbf{R}^2)$ the equality $f(\mathbf{R}^2)=\mathbf{R}^2$ may not hold.
Since every map $f\in \mbox{Emb}(\mathbf{R}^2)$ is open (see \cite{libroembeddings}), we will say that $f$ is a homeomorphism if $f$ is a topological embedding defined onto $\mathbf{R}^2$.
The set of homeomorphisms of the plane will be denoted by $\mbox{Hom}(\mathbf{R}^2)$.
When $\mathcal{H}$ is one of these sets we denote by $\mathcal{H}^ {+}$ (and $\mathcal{H}^ {-}$) the subset of orientation preserving (reversing) elements of $\mathcal{H}$.
We denote by $\mbox{Fix}(f)$ the set of fixed points of a continuous map $f: \mathbf{R}^2 \to \mathbf{R}^2$.
Let $\omega(p)$ be the set of points $q$ for which there is a sequence $n_j\to+\infty$ such that $f^{n_j}(q)\to p$.
If $f\in Hom(\mathbf{R}^2)$ then $\alpha(p)$ denotes the set $\omega(p)$ under $f^{-1}$.
Let $f\in \mbox{Emb}(\mathbf{R}^2)$ and $p\in \mathbf{R}^2$. We say that $\omega(p)=\infty$ if $\norm{f^n(p)}\to \infty$ as $n$ goes to $ \infty$. Analogously, if $f\in \mbox{Hom}(\mathbf{R}^2)$, we say that $\alpha(p)=\infty$ if $\norm{f^{-n}(p)}\to \infty$ as $n$ goes to $ \infty$.
We say that a map $f\in \mbox{Emb}(\mathbf{R}^2)$ is \emph{dissipative} if there exists a compact set $W\subset \mathbf{R}^2$ that is positively invariant and attracts uniformly all compact sets. This means that $f(W)\subset W$ and for each $x\in \mathbf{R}^2$,
$$
dist(f^n(x),W)\to 0, \quad \mbox{ as } \; n\to \infty
$$
uniformly on balls $\abs{x}\leq r$, $r>0$. Observe that in the case $f\in \mbox{Hom}(\mathbf{R}^2)$ the dissipativity of $h$ means that $\infty$ is asymptotically stable for $f^{-1}$.
We say that $0\in \mbox{Fix}(f)$ is a \emph{local attractor} if its basin of attraction ${\mathcal U}=\{p\in \mathbf{R}^ 2 : \omega(p)=\{0\}\}$ contains an open neighbourhood of $0$ in $\mathbf{R}^2$ and that $0$ is a \emph{global attractor} if ${\mathcal U}=\mathbf{R}^2$. The origin is a \emph{stable fixed point} if for every neighborhood $U$ of $0$ there exists another neighborhood $V$ of $0$ such that $f(V)\subset V$ and $f(V)\subset U$. Therefore, the origin is an \emph{asymptotically local (global) attractor} or a \emph{(globally) asymptotically stable fixed point} if it is a stable local (global) attractor. See \cite{Bhatia} for examples.
We say that $0\in \mbox{Fix}(f)$ is a \emph{local repellor} if there exists a neighbourhood $V$ of $0$ such that $\omega(p)\notin V$ for all $0\neq p\in \mathbf{R}^ 2$ and a \emph{global repellor} if this holds for $V=\mathbf{R}^2$.
We say that the origin is an \emph{asymptotically global repellor} if it is a global repellor
and, moreover, if for any neighbourhood $U$ of $0$ there exists another neighbourhood $V$ of $0$, such that,
$V\subset f(V)$ and $V\subset f(U)$.
When the origin is a fixed point of a $C^ 1$-map of the plane we say the origin is a \emph{local saddle} if the two eigenvalues of $Df_0$, $\alpha, \beta$, are both real and verify $0<\abs{\alpha}<1<\abs{\beta}$. In case the two eigenvalues are strictly positive we say the origin is a direct saddle.
We say that the origin is a {\em global (topological) saddle} for a $C^1-$homeomorphism if additionally
its stable and unstable manifolds
$W^{s}(0,f)$, $W^{u}(0,f)$ are unbounded sets that do not accumulate on each other, except at $0$ and $\infty$,
and such that $$\mathbf{R}^2 \setminus (W^{s}\cup W^{u} \cup \{\infty\}) = U_1 \cup U_2 \cup U_3 \cup U_4,$$ where for all $i=1,...,4$ $\; U_i\subset \mathbf{R}^2$ is open connected and homeomorphic to $\mathbf{R}^2$ verifying:
\begin{itemize}
\item[i)] either $f(U_i)=U_i$ or there exists an involution $\varphi:\mathbf{R}^2 \to \mathbf{R}^2$ such that $(f \circ \varphi) (U_i)=U_i$
\item[ii)] for all $p\in U_i$ both $\norm{f^n(p)}\to \infty$ and $\norm{f^{-n}(p)}\to \infty$ as $n$ goes to $\infty$
\end{itemize}
We say that $f\in \mbox{Emb}(\mathbf{R}^2)$ has \emph{trivial dynamics} if $\omega(p) \subset \mbox{Fix}(f)$, for all $p\in \mathbf{R}^ 2$. Moreover, we say that a planar homeomorphism has trivial dynamics if both $\omega(p), \alpha(p) \subset \mbox{Fix}(f)$, for all $p\in \mathbf{R}^ 2$.
Let $f: \mathbf{R}^N \rightarrow \mathbf{R}^N$ be a continuous map.
Let $\gamma : [0,\infty) \to \mathbf{R}^2$ be a topological
embedding of $[0,\infty) \,. \;$ As usual, we identify $\gamma$
with $\gamma\,([0,\infty))\,.$ We will say that $\gamma$ is an \emph{
$f$-invariant ray} if $\, \gamma(0)=(0,0)\,, \,$ $\,f(\gamma)\subset
\gamma \,, \,$ and $\lim_{t\to\infty}|\gamma(t)|=\infty$, where
$|\cdot| \,$ denotes the usual Euclidean norm.
\begin{proposition}[Alarc\'on {\em et al.} \cite{Alarcon}] \label{lemrayo} Let $f\in \mbox{Emb}^ {+}(\mathbf{R}^2)$ be such that $\mbox{Fix}(f)=\{0\}$. If there exists an $f$-invariant ray $\gamma$, then $f$ has trivial dynamics.
\end{proposition}
\begin{corollary} Let $f\in \mbox{Hom}^{+}(\mathbf{R}^2)$ be such that $\mbox{Fix}(f)=\{0\}$. If there exists an $f$-invariant ray $\gamma$, then for each $p\in\mathbf{R}^2$, as $n$ goes to $\pm \infty$, either $f^n(p)$ goes to $0$ or $\norm{f^n(p)}\to \infty$.
\end{corollary}
In order to explain the construction of examples in Section \ref{secDenjoy} we need to introduce the concept of prime end.
We say that $f:\mathbf{R}^2 \to \mathbf{R}^2$ is an \emph{admissible homeomorphism} if
$f$ is orientation preserving, dissipative and has an
asymptotically stable fixed point with proper and unbounded basin
of attraction $U\subset \mathbf{R}^2$. Note that $U$ is non empty, so the
proper condition follows when the fixed point is not a global
attractor. Since $f(U)=U$, we can obtain automatically the
unboundedness condition if we suppose that $f$ is area
contracting.
Let $f:\mathbf{R}^2 \to \mathbf{R}^2$ be an admissible homeomorphism and consider
the compactification of $f$ to the Riemann sphere $f:\mathbf{S}^2 \to \mathbf{S}^2$.
Hence $U\subset \mathbf{S}^2=\mathbf{R}^2 \cup
\{\infty\}$. A \emph{crosscut} $C$ of $U$ is an arc homeomorphic
to the segment $[0,1]$ such that $a,b\notin U$ and $\dot{C}= C
\setminus \{a,b\} \subset U$, where $a$ and $b$ are the extremes of
$C$. Every crosscut divides $U$ into two connected components
homeomorphic to the open disk $\D=\{z\in \mathbf{C} : \abs{z}<1\}$.
Let $x^*$ be a point in $U$. For convenience we will consider only
the crosscut such that $x^* \notin C$. We denote by $D(C)$ the
component of $U\setminus C$ that does not contain $x^*$. A
\emph{null-chain} is a sequence of pairwise disjoint crosscuts
$\{C_n\}_{n\in \mathbf{N}}$ such that $$\lim_{n\to \infty}
\mbox{diam}(C_n)=0 \mbox{ and } D(C_{n+1})\subset D(C_n),$$ where
$\mbox{diam}(C_n)$ is the diameter of $C_n$ on the Riemann sphere.
Two \emph{null-chains} are \emph{equivalent} $\{C_n\}_{n\in \mathbf{N}}
\sim\{C_n^*\}_{n\in \mathbf{N}}$ if given $m\in \mathbf{N}$
$$D(C_n)\subset D(C_m^*) \mbox{ and } D(C_n^*)\subset D(C_m),$$
for $n$ large enough. A \emph{prime end} is defined as a class of
equivalence of a null-chain and the space of prime ends is
$$\mathbf{P}=\mathbf{P}(U)=\mathcal{C}/\sim,$$
where $\mathcal{C}$ is the set of all
null-chains of $U$.
The disjoint union $U^*=U\cup \mathbf{P}$ is a topological space
homeomorphic to the closed disk $\bar{\D}=\{z\in \mathbf{C} :
\abs{z}\leq1\}$ such that its boundary is precisely $\mathbf{P}$.
It is well studied in \cite{pommerenke} that the Theory of Prime
Ends implies that an admissible homeomorphism $f$ induces an
orientation preserving homeomorphism $f^{*}:\mathbf{P} \to
\mathbf{P}$ in the space of prime ends. This topological space is
homeomorphic to the circle, that is $\mathbf{P} \simeq \mathbf{T}$, and
hence the rotation number of $f^{*}$ is well defined, say $\bar
\rho \in \mathbf{T}$. The \emph{rotation number} for an admissible
homeomorphism is defined by $\rho(f)=\bar \rho$.
\subsection{Equivariant Theory}
Let $\Gamma$ be a compact Lie group acting on $\mathbf{R}^2$, that is, a group which has the structure of a compact $C^{\infty}$-differentiable manifold such that the map $\Gamma \times \Gamma \to \Gamma$, $(x,y)\mapsto xy^{-1}$ is of class $C^{\infty}$. The folllowing definitions and results are taken from Golubitsky {\em et al.} \cite{golu2}, especially Chapter~XII, to which we refer the reader interested in further detail.
We think of a group mostly through its action or representation on $\mathbf{R}^2$.
A {\em linear action} of $\Gamma$ on $\mathbf{R}^2$ is a continuous mapping
\begin{eqnarray*}
\Gamma \times \mathbf{R}^2 & \rightarrow & \mathbf{R}^2 \\
(\gamma , p ) & \mapsto & \gamma p
\end{eqnarray*}
such that, for each $\gamma \in \Gamma$ the mapping $\rho_{\gamma}$ that takes $p$ to $\gamma p$ is linear and, given $\gamma_1, \gamma_2 \in \Gamma$, we have $\gamma_1(\gamma_2 p)= (\gamma_1 \gamma_2) p$.
Furthermore the identity in $\Gamma$ fixes every point. The mapping $\gamma\mapsto \rho_{\gamma}$ is called the {\em representation} of $\Gamma$ and describes how each element of $\Gamma$ transforms the plane.
We consider only standard group actions and representations.
A representation of a group $\Gamma $ on a vector space $V$ is {\em absolutely irreducible} if the only linear mappings on $V$ that commute with $\Gamma $ are scalar multiples of the identity.
Given a map $f:\mathbf{R}^2\longrightarrow\mathbf{R}^2$,
we say that $\gamma \in \Gamma$ is a \emph{symmetry} of $f$ if $f(\gamma x)=\gamma f(x)$.
We define the \emph{symmetry group} of $f$ as the biggest closed subset of $GL(2)$ containing all the symmetries of $f$. It will be denoted by $\Gamma_f$.
We say that $f:\mathbf{R}^2\to \mathbf{R}^2$ is \emph{$\Gamma$-equivariant} or that $f$ {\em commutes} with $\Gamma$ if
$$
f(\gamma x)=\gamma f(x) \quad \mbox{ for all }\quad \gamma \in \Gamma.
$$
It follows that
every map $f:\mathbf{R}^2\to \mathbf{R}^2$ is equivariant under the action of its symmetry group, that is, $f$ is $\Gamma_f$-equivariant.
Let $\Sigma$ be a subgroup of $\Gamma$. The {\em fixed-point subspace} of $\Sigma$ is
$$
\mbox{Fix} (\Sigma) =\{p\in \mathbf{R}^2: \sigma p=p \; \mbox{ for all } \; \sigma \in \Sigma\}.
$$ If $\Sigma$ is generated by a single element $\sigma \in \Gamma$, we write \emph{$\mbox{Fix}\langle\sigma\rangle$}.
We note that, for each subgroup $\Sigma$ of $\Gamma$, $\mbox{Fix} (\Sigma)$ is invariant by the dynamics of a $\Gamma$-equivariant map (\cite{golu2}, XIII, Lemma 2.1).
When $f$ is $\Gamma$-equivariant, we can use the symmetry to generalize information obtained for a particular point. This is achieved through the {\em group orbit} $\Gamma x$ of a point $x$, which is defined to be
$$
\Gamma x = \{ \gamma x: \; \; \gamma \in \Gamma\}.
$$
\begin{lemma}
Let $f:\mathbf{R}^2\to \mathbf{R}^2$ be $\Gamma$-equivariant and let $p$ be a fixed point of $f$. Then all points in the group orbit of $p$ are fixed points of $f$.
\end{lemma}
\begin{proof}
If $f(p)=p$ it follows that $f(\gamma p) = \gamma f(p) = \gamma p$, showing that $\gamma p$ is a fixed point of $f$ for all $\gamma \in \Gamma$.
\end{proof}
The relation between the group action and the Jacobian matrix of an equivariant map $f$ is obtained through the following
\begin{lemma}\label{LemmaJacobian}
Let $f: \; V \rightarrow V$ be a $\Gamma$-equivariant map differentiable at the origin. Then $Df(0)$, the Jacobian of $f$ at the origin, commutes with $\Gamma$.
\end{lemma}
\begin{proof}
Since $f$ is $\Gamma $-equivariant we have
$f(\gamma .v ) = \gamma f(v)$ for all $\gamma \in \Gamma$ and $v \in V$.
Differentiating both sides of the equality with respect to $v$, we obtain
$
Df(\gamma . v) \gamma = \gamma Df(v)
$
and, evaluating at the origin gives
$
Df(0) \gamma = \gamma Df(0).
$
\end{proof}
\section{Symmetries in the plane}
In this section, we describe the consequences for the local dynamics arising from the fact that a map is equivariant under the action of a compact Lie group $\Gamma$. These are patent in the structure of the Jacobian matrix at the origin, obtained using Lemma~\ref{LemmaJacobian}.
Since every compact Lie group in $GL(2)$ can be identified with a subgroup of the orthogonal group $O(2)$, we need only be concerned with the groups we list below.
\paragraph{Compact subgroups of $O(2)$}
\begin{itemize}
\item $O(2)$, acting on $\mathbf{R}^2 \simeq \mathbf{C}$ as the group generated by $\theta$ and $\kappa$ given by
$$
\theta . z = e^{i\theta} z, \quad \theta \in S^1 \qquad\mbox{ and } \qquad \kappa . z=\bar{z}.
$$
\item $SO(2)$, acting on $\mathbf{R}^2 \simeq \mathbf{C}$ as the group generated by $\theta$ given by
$$
\theta . z = e^{i\theta} z, \quad \theta \in S^1.
$$
\item $D_n$, $n \geq 2$, acting on $\mathbf{R}^2 \simeq \mathbf{C}$ as the finite group generated by $\zeta$ and $\kappa$ given by
$$
\zeta . z = e^{\frac{2\pi i}{n}} z \qquad\mbox{ and } \qquad \kappa . z=\bar{z}.
$$
\item $\mathbf{Z}_n$, $n \geq 2$, acting on $\mathbf{R}^2 \simeq \mathbf{C}$ as the finite group generated by $\zeta$ given by
$$
\zeta . z = e^{\frac{2\pi i}{n}} z.
$$
\item $\mathbf{Z}_2(\langle\kappa\rangle)$, acting on $\mathbf{R}^2$ as
$$
\kappa . (x,y) = (x, -y).
$$
\end{itemize}
Since most of our results depend on the existence of a unique fixed point for $f$,
it is worthwhile noting that
the group actions we are concerned with are such that $\mbox{Fix} (\Gamma) = \{ 0 \}$.
Therefore, if $f$ is $\Gamma$-equivariant then $f(0)=0$.
If the representation is absolutely irreducible, we know that $Df(0)$ is a multiple of the identity and thus it has one real eigenvalue of geometric multiplicity two. Therefore, the origin is locally either an attractor or a repellor.
We have the following
\begin{lemma}
The standard representation on $\mathbf{R}^2$ is absolutely irreducible for $O(2)$ and $D_n$ with $n \geq 3$ and for no other subgroup of $O(2)$.
\end{lemma}
\begin{proof}
The proof follows by direct computation.
\begin{itemize}
\item $O(2)$: the generators of this group are $\theta$ and $\kappa$ and it suffices to find the linear matrices that commute with both. A real matrix
$$
\left(\begin{array}{cc}
a & b \\
c & d
\end{array} \right)
$$
commutes with $\kappa$ if and only if $b=c=0$. In order for such a matrix to commute with any rotation it must be
$$
\left( \begin{array}{cc}
a & 0 \\
0 & d
\end{array} \right) \left(\begin{array}{cc}
\cos{\theta} & -\sin{\theta} \\
\sin{\theta} & \cos{\theta}
\end{array} \right) = \left(\begin{array}{cc}
\cos{\theta} & -\sin{\theta} \\
\sin{\theta} & \cos{\theta}
\end{array} \right) \left(\begin{array}{cc}
a & 0 \\
0 & d
\end{array} \right)
$$
which holds when $a=d$ or $\sin{\theta}=0$ for all $\theta \in S^1$.
Therefore, the action of $O(2)$ is absolutely irreducible.
\item $SO(2)$: the elements of $SO(2)$ are rotation matrices which commute with any other rotation matrix, also non-diagonal ones.
\item $D_n$, $n \geq 3$: see the proof for $O(2)$. In the last step, we must have $a=d$ or $\sin{2\pi i/n} = 0$ which is never satisfied for $n \geq 3$. Hence, the action is absolutely irreducible.
\item $\mathbf{Z}_n$, $n \geq 3$: as for $SO(2)$, any rotation matrix commutes with the rotation of $2\pi /n$, including non-diagonal ones.
\item $\mathbf{Z}_2(\langle\kappa\rangle)$: see the proof for $\kappa \in O(2)$ to conclude that linear commuting matrices are diagonal but not necessarily linear multiples of the identity.
\item $\mathbf{Z}_2$: all linear maps commute with $-Id$.
\item $D_2=\mathbf{Z}_2 \oplus \mathbf{Z}_2(\langle\kappa\rangle)$: as above, $\mathbf{Z}_2$ introduces no restrictions and for commuting with $\kappa$ it suffices that the map is diagonal.
\end{itemize}
\end{proof}
The following result is then a straightforward consequence of the previous proof.
\begin{lemma}
The linear maps that commute with the standard representations of the subgroups of $O(2)$ are rotations and homotheties (and their compositions) for $SO(2)$ and $\mathbf{Z}_n$, $n \geq 3$, linear multiples of the identity for $O(2)$ and $D_n$, $n \geq 3$, any linear map for $\mathbf{Z}_2$ and maps represented by diagonal matrices for the remaining groups.
\end{lemma}
\begin{proof}
The only linear maps that were not already explicitly calculated in the previous proof are those that commute with rotations. We have
$$
\left( \begin{array}{cc}
a & b \\
c & d
\end{array} \right) \left(\begin{array}{cc}
\cos{\theta} & -\sin{\theta} \\
\sin{\theta} & \cos{\theta}
\end{array} \right) = \left(\begin{array}{cc}
\cos{\theta} & -\sin{\theta} \\
\sin{\theta} & \cos{\theta}
\end{array} \right) \left(\begin{array}{cc}
a & b \\
c & d
\end{array} \right)
$$
if and only if either $\sin{\theta}=0$ for all $\theta \in S^1$ or $a=d$ and $b=-c$. Hence, the only maps commuting with either $SO(2)$ or $\mathbf{Z}_n$, $n \geq 3$, are rotations and homotheties and their compositions.
\end{proof}
With the results obtained so far, we are able to describe the Jacobian matrix at the origin for maps equivariant under each of the groups above.
\begin{proposition}[Proposition 2.3 in \cite{SofisbeGlobal}]
Let $f$ be a planar map differentiable at the origin. The admissible forms for the Jacobian matrix of $f$ at the origin are those given in Table \ref{Jacobian} depending on the symmetry group of $f$.
\end{proposition}
\begin{table}
\begin{center}
\begin{tabular}{ccr}
\hline {}&{}&{}\\
Symmetry group & $Df(0)$ & hyperbolic local dynamics \\ {}&{}&{}\\
\hline {}&{}&{}\\
$O(2)$ & $\left( \begin{array}{cc} \alpha & 0 \\ 0 & \alpha \end{array} \right)$ $\alpha \in \mathbf{R}$
&attractor / repellor \\ {}&{}&{}\\
\hline {}&{}&{}\\
$SO(2)$ & $\left( \begin{array}{cc} \alpha & -\beta \\ \beta & \alpha \end{array} \right)$ $\alpha, \beta \in \mathbf{R}$ &attractor / repellor \\ {}&{}&{}\\
\hline {}&{}&{}\\
$D_n, \;n\geq 3$ & $\left( \begin{array}{cc} \alpha & 0 \\ 0 & \alpha \end{array} \right)$ $\alpha \in \mathbf{R}$
&attractor / repellor \\ {}&{}&{}\\
\hline {}&{}&{}\\
$\mathbf{Z}_n, \;n\geq 3$ & $\left( \begin{array}{cc} \alpha & -\beta \\ \beta & \alpha \end{array} \right)$ $\alpha, \beta \in \mathbf{R}$ &attractor / repellor \\ {}&{}&{}\\
\hline {}&{}&{}\\
$\mathbf{Z}_{2}$ & any matrix &saddle / attractor / repellor \\ {}&{}&{}\\
\hline {}&{}&{}\\
$\mathbf{Z}_2(\langle\kappa\rangle)$ & $\left( \begin{array}{cc} \alpha & 0 \\ 0 & \beta \end{array} \right)$ $\alpha, \beta \in \mathbf{R}$ &saddle / attractor / repellor \\ {}&{}&{}\\
\hline {}&{}&{}\\
$D_{2}=\mathbf{Z}_2 \oplus \mathbf{Z}_{2}(\langle\kappa\rangle)$\quad & \quad $\left( \begin{array}{cc} \alpha & 0 \\ 0 & \beta \end{array} \right)$ $\alpha, \beta \in \mathbf{R}$ \quad &saddle / attractor / repellor \\ {}&{}&{}\\
\hline
\end{tabular}
\end{center}
\caption{Compact subgroups of $O(2)$ and the admissible forms of the Jacobian at the origin of maps equivariant under the standard action of each group.
If in addition the Jacobian at the origin is hyperbolic, then this determines the local stability.}\label{Jacobian}
\end{table}
Furthermore, the symmetry constrains the normal form as described in \cite[Theorem 2.1]{SofisbePolynomial}
and in the next result, and its consequences for the linear part of $f$ appear in Table~\ref{Jacobian}.
\begin{proposition}[Proposition 3.1 in \cite{SofisbePolynomial}]
Let $\Gamma$ be a compact Lie group acting on $\mathbf{R}^2$. Assume $\Gamma$ is the symmetry group of a polynomial map $f$.
\begin{itemize}
\item[(i)] If $\kappa \in \Gamma$, then $f$ does not answer the DMYQ($2$) in the affirmative unless $f$ is of the form:
$$
f(x,y)=\left(\begin{array}{cc}d_1&0\\0&d_2\end{array}\right)\left(\begin{array}{c}x\\ y\end{array}\right)+
y^2p(y^2)\left(\begin{array}{c}1\\ 0\end{array}\right)\ .
$$
\item[(ii)] If there is an element $\zeta \in \Gamma$ of order $n \geq 3$, then $f$ does not answer the DMYQ($2$) in the affirmative unless $f$ is linear.
Moreover, the linear part of $f$ is either a homothety or a rotation matrix.
\end{itemize}
\end{proposition}
\section{Dynamics --- local to global}
\begin{figure}\begin{center}
a)\includegraphics[scale=0.40]{1-O2-poster.eps}\qquad
b)\includegraphics[scale=0.40]{3-D4-posterC.eps}\qquad
c)\includegraphics[scale=0.40]{4-D2-posterC.eps}\qquad
d)\includegraphics[scale=0.40]{6-Z2k-posterC.eps}
\end{center}
\caption{
Local/global attractor with symmetry:
a) $O(2)$;
b) $D_4$ (without symmetries $\mathbf{Z}_8$ or $SO(2)$);
c) $D_2$ (without symmetry $D_4$);
d) $\mathbf{Z}_2\langle\kappa\rangle$ (without symmetry $D_4$)
\label{ComFlip}}\end{figure}
Figure~\ref{ComFlip} illustrates the dynamics near the origin of equivariant maps for several symmetry groups.
A common feature of Figures~\ref{ComFlip} a)--d) is the existence of at least one symmetry axis.
This axis is the subspace fixed by a reflection and hence it is invariant under the dynamics.
Such a fixed-point subspace naturally contains an invariant ray
(see \cite[Lemma 3.3]{SofisbeGlobal}).
This allows us to use Proposition~\ref{lemrayo} to obtain the following results:
\begin{proposition}[Proposition 3.4 in \cite{SofisbeGlobal}] \label{proprayoS} Let $f\in \mbox{Emb}(\mathbf{R}^2)$ have symmetry group $\Gamma$ with $\kappa \in \Gamma$, such that $\mbox{Fix}(f)=\{0\}$. Suppose one of the following holds:
\begin{itemize}
\item[$a)$] $f\in \mbox{Emb}^ {+}(\mathbf{R}^2)$ and $f$ does not interchange connected components of $\mathbf{R}^2 \setminus \mbox{Fix} \langle\kappa\rangle$.
\item[$b)$] $\mbox{Fix}(f^ 2)=\{0\}$.
\end{itemize}
Then for each $p\in \mathbf{R}^2$ either $\omega(p)=\{0\}$ or $\omega(p)=\infty$.
\end{proposition}
The next example shows that assumption $b)$ in Proposition \ref{proprayoS} is necessary in the case where $f$ interchanges connected components of $\mathbf{R}^2 \setminus \mbox{Fix} \langle\kappa\rangle$.
\paragraph{Example:} Consider the map $f: \mathbf{R}^2 \to \mathbf{R}^2$ defined by
$$
f(x,y)=\left(-ax^3+(a-1)x,-\frac{y}{2}\right) \quad 0<a<1.
$$
It is easily checked that $f$ has symmetry group $D_2$ and verifies (see Figure \ref{figureattractor}):
\begin{enumerate}
\item $f\in \mbox{Emb}^{+}(\mathbf{R}^2)$ is an orientation-preserving diffeomorphism.
\item $\mbox{Spec}(f)\cap [0, \infty)=\emptyset$.
\item $0$ is a local hyperbolic attractor.
\item $\mbox{Fix}(f^2)\neq \{0\}$.
\end{enumerate}
\medskip
\begin{figure}[htp]
\sidecaption
\includegraphics[width=2in]{period_2_point.eps}\\
\caption{A local attractor which is not a global attractor due to the existence of periodic orbits.}\label{figureattractor}
\end{figure}
\begin{theorem}[Theorem 3.5 in \cite{SofisbeGlobal}] \label{proprayoS1} Let $f\in \mbox{Emb}(\mathbf{R}^2)$ be dissipative with symmetry group $\Gamma$ with $\kappa \in \Gamma$ such that $\mbox{Fix}(f)=\{0\}$. Suppose in addition that one of the following holds:
\begin{itemize}
\item[$a)$] $f\in \mbox{Emb}^ {+}(\mathbf{R}^2)$ and $f$ does not interchange connected components of $\mathbf{R}^2 \setminus \mbox{Fix} \langle\kappa\rangle$.
\item[$b)$] There exist no $2$-periodic orbits.
\end{itemize}
Then $0$ is a global attractor.
\end{theorem}
\begin{corollary} [Corollary 3.6 in \cite{SofisbeGlobal}]Suppose the assumptions of Theorem \ref{proprayoS1}
are verified and $f$ is differentiable at $0$. If every eigenvalue of $Df(0)$ has norm strictly less than one, then $0$ is a global asymptotic attractor.
\end{corollary}
For analogous results concerning a repellor see \cite{SofisbeGlobal}.
\begin{figure}
\sidecaption
a) \includegraphics[scale=0.35]{5-D2-posterC.eps}
\quad
b) \includegraphics[scale=0.35]{7-Z2k-posterC.eps}
\quad
c) \includegraphics[scale=0.35]{9-Z2-poster.eps}
\caption{
Local/global saddle with symmetry:
a) $D_2$;
b) $\mathbf{Z}_2(\kappa)$;
c) $\mathbf{Z}_2$.
\label{figSaddle}}\end{figure}
For the groups $\mathbf{Z}_{2}$, $\mathbf{Z}_2(\langle\kappa\rangle)$ and $D_{2}=\mathbf{Z}_2 \oplus \mathbf{Z}_{2}(\langle\kappa\rangle)$
the origin may also be a saddle as illustrated in Figure~\ref{figSaddle}.
For $D_{2}$, we have:
\begin{proposition} [\cite{SofisbeSaddle}]
Let $f\in \mbox{Hom}(\mathbf{R}^ 2)$ be a $C^1$-homeomorphism with symmetry group $D_2$ such that $Fix(f)=\{0\}$. Suppose also that one of the following holds:
\begin{itemize}
\item[$a)$] $f$ is orientation preserving and $0$ is a direct saddle.
\item[$b)$] There exist no $2$-periodic orbits.
\end{itemize}
Then if $0$ is a local saddle, then $0$ is a global saddle.
\end{proposition}
In order to obtain a global saddle for $f$ with symmetry group either $\mathbf{Z}_{2}$ or $\mathbf{Z}_2(\langle\kappa\rangle)$,
we need the additional assumption that $f$ is a diffeomorphism, see \cite{SofisbeSaddle}.
\section{Strictly Local Dynamics} \label{secDenjoy}
Figure~\ref{Z4local} shows the local dynamics for maps equivariant under the action of groups
that do not contain a reflection. These are $SO(2)$ and $\mathbf{Z}_n$. For these groups, local dynamics of attractor/repellor type does not necessarily extend to global dynamics, as we proceed to indicate.
We use examples referring to a local attractor, examples with a local repellor may be obtained considering $f^{-1}$.
\begin{figure}[hhh]
\sidecaption
a) \includegraphics[scale=0.25]{8-Z2-posterB.eps}
\qquad
b) \includegraphics[scale=0.25]{10-Z4-posterB.eps}
\caption{
Local attractor with symmetry: a) $\mathbf{Z}_2$; b) $\mathbf{Z}_4$.
\label{Z4local}}
\end{figure}
\bigbreak
The dynamics of an $SO(2)$-symmetric embedding is mostly determined by its radial component,
as can be seen by writing $f$ in polar coordinates as $f(\rho, \theta)=(R(\rho, \theta), T(\rho, \theta))$.
It is easily shown that since $f$ is $SO(2)$-equivariant, the radial component $R(\rho, \theta)$ only depends on $\rho$ and $R\in Emb(\mathbf{R}^+)$.
The fixed points of the radial component are invariant circles for $f$ hence knowledge about local dynamics does not contribute to the description of global dynamics unless $\mbox{Fix}(R)= \{0\}$.
\bigbreak
The next two theorems show how a local attractor may be prevented from being a global attractor in a $\mathbf{Z}_n$-equivariant problem. Thus the examples in Figures~ \ref{Z4local} a) and b) may or may not extend to the whole plane.
\begin{theorem}[Theorem 3.1 in \cite{SofisbeSzlenk}] \label{exper} For each $n\ge 2$ there exists $f:\mathbf{R}^2\to\mathbf{R}^2$ such
that:
\begin{enumerate}
\renewcommand{\theenumi}{\alph{enumi}}
\renewcommand{\labelenumi}{{\theenumi})}
\item\label{C1a} $f$ is a differentiable homeomorphism;
\item\label{C1b} $f$ has symmetry group $\mathbf{Z}_n$;
\item \label{C1E} $Fix(f)=\{0\}$;
\item\label{C1d} The origin is a local attractor;
\item \label{C3E} There exists a periodic orbit of minimal period $n$.
\end{enumerate}
\end{theorem}
The idea of the proof is to start with a $\mathbf{Z}_4$-equivariant example due to Szlenk (see \cite{Cima-Manosa}), sharing the same properties.
Each quadrant of the plane is invariant under the map $f_4$ of this example.
We deform the first quadrant into a sector of the plane, of angle $2\pi/n$ and then use the $\mathbf{Z}_n$ symmetry to cover the rest of the plane, as illustrated in Figure~\ref{FigSlzenkZn}.
The main difficulty is to prove that the result is a differentiable homeomorphism.
\begin{figure}
\sidecaption
\includegraphics[scale=.50] {Fn.eps}
\caption{Construction of the $\mathbf{Z}_n$-equivariant example $F_n$ in a fundamental domain of the
$\mathbf{Z}_n$-action, shown here for $n=6$.\label{FigSlzenkZn}}
\end{figure}
The $\mathbf{Z}_m$-equivariant homeomorphisms constructed in Theorem \ref{exper} have rotation number
$1/m$. So we might be led to think that the presence of the $\mathbf{Z}_m$-symmetry
implies that the rotation number of the homeomorphism should be
rational. One consequence would be that the asymptotically stable fixed point
is a global attractor if and only if there are no periodic points different from the fixed point.
The next result shows that this is false. We prove in \cite{AlarconDenjoy} the existence of $\mathbf{Z}_m$-equivariant and dissipative homeomorphisms in the plane with an asymptotically stable fixed point such that the induced map in the space of prime ends is conjugated to a
Denjoy map, which is also $\mathbf{Z}_m$-equivariant.
The idea is to construct $\mathbf{Z}_m$-equivariant Denjoy maps in the circle and then, in
the context of symmetry, to reproduce the construction used to prove the following:
\begin{proposition}[Proposition 2 in \cite{irrationalRotNumer}] Given $w \in (0,1)\setminus \mathbf{Q}$ and a Denjoy map $\phi$, there exists an admissible map $f$ with rotation number $\bar{w}$ and such that $f^{*}$
is topologically conjugated to $\phi$.
\end{proposition}
Observe that two admissible homeomorphisms $f_1, f_2$ with the same basin of
attraction $U$ verify that $$(f_1 \circ f_2)^*=f_1^*\circ f_2^*.$$
Let $f$ be an admissible homeomorphisms with basin of attraction U.
Suppose $f$ is $Z_m$-equivariant and $U$ is also invariant by $R_{\frac{1}{m}}$. Hence, the following holds: $$f^*\circ
R_{\frac{1}{m}}^*=R_{\frac{1}{m}}^*\circ f^*.$$
Since $R_{\frac{1}{m}}^*$ is a periodic homeomorphism of
$\mathbf{T}^1$ with rotation number $1/m$, then
$R_{\frac{1}{m}}^*$ is conjugated to the linear rotation
$R_{\frac{1}{m}}$ and $f^*$ is said to be $\mathbf{Z}_m$-\emph{equivariant in
the space of prime ends}.
\begin{theorem}[Theorem 4.2 in \cite{AlarconDenjoy}] \label{teoDenjoyZm} Given an irrational number $\tau\notin \mathbf{Q}$,
there exists a $\mathbf{Z}_m-$equivariant and admissible homeomorphism in
$\mathbf{R}^2$ with rotation number $\bar \tau \in \mathbf{T}$ and such that
induces a Denjoy map in the circle of prime ends which is also
$\mathbf{Z}_m-$equivariant.
\end{theorem}
Hence, \cite{AlarconDenjoy} shows that for $\mathbf{Z}_m$-equivariant homeomorphisms one cannot guarantee that the rotation number is rational and proves the existence of $\mathbf{Z}_m$-equivariant homeomorphisms
with some complicated and interesting dynamical features.
\begin{acknowledgement}The research of all authors at Centro de Matem\'atica da Universidade do Porto (CMUP)
had financial support from
the European Regional Development Fund through the programme COMPETE and
from the Portuguese Government through the Funda\c c\~ao para
a Ci\^encia e a Tecnologia (FCT) under the project PEst-C/MAT/UI0144/2011.
B.\ Alarc\'{o}n was also supported by grant MICINN-12-MTM2011-22956 of the Ministerio de Ciencia e
Innovaci\'on (Spain).
\end{acknowledgement}
|
\section{Nonequilibrium bosonic dynamical mean field theory}
\label{app:BDMFT}
The bosonic dynamical mean field theory (BDMFT) for the Bose-Hubbard model in \emph{equilibrium} has been derived in Ref.~\cite{Anders:2011uq} in three alternative ways, using the kinetic energy functional, an effective medium approach, and the quantum cavity method, which is very similar to the cavity construction by Snoek and Hofstetter \cite{Snoek:2010uq}.
In this appendix we follow the latter approach, which combines the cavity construction with a generating functional formalism and a cumulant expansion to second order. By performing the derivation on the three branch Kadanoff-Baym contour $\mathcal{C}$ we obtain the \emph{nonequilibrium} generalization of BDMFT.
We will also show that BDMFT corresponds to the first order correction in the inverse coordination number $1/z$, and the second order correction in the fluctuations, of the mean-field approximation for the Bose-Hubbard model.
\subsection{The Bose-Hubbard model}
We consider the Bose-Hubbard model [Eq.\ (\ref{eq:Model})]
\begin{equation*}
H = -J \sum_{\langle i, j \rangle} ( b^\dagger_ib_j + b^\dagger_jb_i )
+\frac{U}{2} \sum_i \hat{n}_i(\hat{n}_i-1) - \mu \sum_i \hat{n}_i \, ,
\end{equation*}
on a lattice with nearest neighbor hopping $J$ and a local pair interaction $U$, where $\hat{n}_i(\hat{n}_i-1) = b^\dagger_i b^\dagger_i b_i b_i$ is a pure two particle interaction counting the number of pairs on site $i$, and $\langle i, j \rangle$ denotes the sum over all nearest neighbor pairs $i$ and $j$.
Using the Nambu-spinor notation $\mbf{b}^{\dagger} = ( b^\dagger, \, b )$ and collecting the local terms on site $i$ into $H_i = U \hat{n}_i(\hat{n}_i-1)/2 - \mu \hat{n}_i$, the Hamiltonian $H$ can be expressed as
\begin{equation}
H = \sum_i H_i - J \sum_{\langle i, j \rangle} \mbf{b}^{\dagger}_i \mbf{b}^{\phantom{\dagger}}_j \, ,
\end{equation}
where we have used that $b_i b^\dagger_j = b^\dagger_j b_i$ if $i \ne j$. Note that in the Nambu notation $\mbf{b}^{\dagger}_i \mbf{b}^{\phantom{\dagger}}_j$ is hermitian, i.e.\
\begin{equation}
\mbf{b}^{\dagger}_i \mbf{b}^{\phantom{\dagger}}_j = \mbf{b}^{\dagger}_j \mbf{b}^{\phantom{\dagger}}_i \, , \quad \textrm{for} \, i \ne j \, .
\label{eq:NambuCommutation}
\end{equation}
\subsection{Kadanoff-Baym and Nambu formalism }
To treat an arbitrary time evolution starting from a finite temperature equilibrium state we formulate the theory on the three-branch Kadanoff-Baym contour $\mathcal{C}$ ($0 \! \rightarrow \! t_\text{max} \! \rightarrow \! 0 \! \rightarrow \! -i\beta$) \cite{Aoki:2014kx, Stefanucci:2013oq}. The partition function $\mathcal{Z}$ of the initial state can be expressed as $\mathcal{Z} = \textrm{Tr}[ \mathcal{T}_\C e^{-iS} ]$ where $S$ is the action defined on the contour $\mathcal{C}$, $S = \int_\mathcal{C} dz \, H(z)$, $\mathcal{T}_\mathcal{C}$ is the time-ordering operator on $\mathcal{C}$, and the trace $\textrm{Tr}[\cdot]$ runs over the Hilbert space of $H$.
Time-dependent operator expectation values can be expressed in the trace formalism as
\begin{equation}
\langle \hat{O}(t) \rangle_S = \frac{1}{\mathcal{Z}} \textrm{Tr}[ \mathcal{T}_\C e^{-iS} \hat{O}(t) ] \, ,
\end{equation}
and the single-particle Green's function on the contour, $G(t, t')$, is given by $G(t, t') = -i \langle b(t) b^\dagger(t') \rangle_S$. The Nambu generalization of the single-particle Green's function is a $2 \times 2$ matrix, which can be expressed in spinor notation as
\begin{equation}
\mbf{G}(t, t') = -i \langle \mbf{b}(t) \mbf{b}^\dagger(t') \rangle \, . \label{eq:NambuGf}
\end{equation}
For a general introduction to the Kadanoff-Baym contour formalism, see Ref.\ \cite{Stefanucci:2013oq} and for a DMFT specific introduction see Ref.\ \cite{Aoki:2014kx}.
\subsection{Real-time generating functional}
To construct the generating functional on the contour $\mathcal{C}$ we introduce source fields $\eta_i$ on each site $i$ and the source action
\begin{equation}
S_\eta = \int_\mathcal{C} dt \, H_\eta(t) \, , \quad \textrm{where }
H_\eta = \sum_i \mbf{b}^{\dagger}_i \boldsymbol{\eta}^{\phantom{\dagger}}_i \, .
\end{equation}
Using $S_\eta$ and the action $S$ of the system
\begin{multline}
S = \int_\mathcal{C} dt \sum_i H_i(t) - J \int_\mathcal{C} dt \sum_{\langle i,j \rangle} \mbf{b}^{\dagger}_i(t) \mbf{b}^{\phantom{\dagger}}_j(t)
\, ,
\end{multline}
the generating functional $\mathcal{Z}[\eta]$ can be defined as
\begin{equation}
\mathcal{Z}[\eta] = \textrm{Tr} \left[ \mathcal{T}_\C \exp[-iS + S_\eta] \right] \, .
\end{equation}
It can be used to compute any \emph{connected} response function by taking derivatives with respect to the source fields
\begin{equation}
\frac{\partial^n}{\partial \eta^{\dagger}_{\alpha_1} \dots \partial \eta^{\dagger}_{\alpha_n}}
\ln \mathcal{Z}[\eta] |_{\eta = 0} =
\langle b_{\alpha_n} \dots b_{\alpha_1} \rangle_S^{(c)}
\, .
\label{eq:ZResponseFunction}
\end{equation}
\subsection{Cavity construction}
To derive a local effective action we use the standard cavity construction \cite{Georges:1996aa} and separate the Hamiltonian into three parts,
\begin{equation}
H = H_0 + \Delta H + H^{(0)} \, ,
\end{equation}
where $H_0$ acts on the site $i=0$, $\Delta H$ connects the zeroth site to its neighbors, and $H^{(0)}$ is the lattice with a cavity at the zeroth site, i.e.\
\begin{align}
H_0 & = -\mu n_0 + \frac{U}{2} n_0 (n_0 - 1) \, , \\
\Delta H & = - J \sum_{\langle 0,i \rangle} \mbf{b}^{\dagger}_i \mbf{b}^{\phantom{\dagger}}_0 \, , \\
H^{(0)} & = \sum_{i\ne 0} H_i
- J \sum_{\begin{subarray}{c} \langle i,j \rangle\\ i,j \ne 0\end{subarray}}
\mbf{b}^{\dagger}_i \mbf{b}^{\phantom{\dagger}}_j \, ,
\end{align}
which in turn separates the action $S$ into
\begin{equation}
S = S_0 + \Delta S + S^{(0)} \, .
\end{equation}
Analogously the source term can be decomposed into
\begin{equation}
H_\eta = H_{0, \eta} + H^{(0)}_\eta , \,
\end{equation}
according to the same protocol, with
\begin{equation}
H_{0, \eta} = \mbf{b}^{\dagger}_0 \boldsymbol{\eta}^{\phantom{\dagger}}_0 , \quad
H^{(0)}_\eta = \sum_{i \ne 0} \mbf{b}^{\dagger}_i \boldsymbol{\eta}^{\phantom{\dagger}}_i \, ,
\end{equation}
which yields the corresponding terms of the source action
\begin{equation}
S_\eta = S_{0, \eta} + S^{(0)}_\eta \, .
\end{equation}
Using this separation of the zeroth site's degrees of freedom the generating functional can be written as
\begin{equation}
\mathcal{Z}[\eta] = \textrm{Tr}_0 \left[ \mathcal{T}_\C e^{-iS_0 + S_{0, \eta}}
\mathcal{Z}^{(0)} \langle e^{-i \Delta S + S^{(0)}_\eta} \rangle_{S^{(0)}}
\right] \, , \label{eq:GeneratingFunctional}
\end{equation}
where $\textrm{Tr}_0[\cdot]$ denotes the trace over the Fock-space of the zeroth site.
In this form the generating functional can be approximated and/or taken to e.g.\ the infinite connectivity limit, which results in different types of dynamical mean field theory (DMFT) approximations.
\subsection{Cumulant expansion}
We are now ready to perform a \emph{cumulant expansion} \cite{JPSJ.17.1100} of the expectation value $\langle e^{-i \Delta S + S^{(0)}_\eta} \rangle_{S^{(0)}}$ in Eq.\ (\ref{eq:GeneratingFunctional}). Formally this corresponds to expanding $\ln \langle e^{-i \Delta S + S^{(0)}_\eta} \rangle_{S^{(0)}}$ in an infinite sum of response functions with respect to $S^{(0)}$. The initial logarithm ensures that the series enters in the exponent, and for this reason the procedure is often referred to as ``re-exponentiation''.
Following Ref.\ \onlinecite{JPSJ.17.1100} the cumulant expansion becomes
\begin{multline*}
\ln \langle \exp[-i \Delta S + S^{(0)}_\eta] \rangle_{S^{(0)}} =
\langle \exp[-i \Delta S + S^{(0)}_\eta] - 1 \rangle_{S^{(0)}}^{(c)}
\\ = \!
\sum_{n=1}^\infty \frac{1}{n!}
\! \int_\mathcal{C} \!\! dt_1 \dots \!\! \int_\mathcal{C} \!\! dt_n
\left\langle \prod_{k=1}^n
( -i \Delta H(t_k) + H^{(0)}_\eta (t_k) )
\right\rangle^{\!\! (c)}_{\!\! S^{(0)}}
\! \! \! \! . \label{eq:CumulantExpansion}
\end{multline*}
In the derivation of the fermionic dynamical mean field effective action, the cumulant expansion terminates at second order in the limit of infinite dimensions $z \rightarrow \infty$ (using a $J \rightarrow J/\sqrt{z}$ scaling of the hopping) \cite{Georges:1996aa}. This yields the usual hybridization function term
\begin{equation*}
\ln \langle e^{-i \Delta S + S^{(0)}_\eta} \rangle_{S^{(0)}} = \dots =
\iint_\mathcal{C} dt \, dt' \, b^\dagger(t) \Delta(t, t') b(t') \, .
\end{equation*}
For Bosons, however, anomalous contributions due to symmetry breaking scale linearly with the coordination number $z$, requiring a $1/z$ scaling of the hopping to obtain a finite $z \rightarrow \infty$ limit \cite{Anders:2010uq, Anders:2011uq}. This procedure results in the mean field effective action \cite{Sachdev:1999fk} which does not include quantum fluctuations of non-condensed Bosons.
In order to retain fluctuations we therefore avoid taking the infinite connectivity limit and instead {\it truncate} the cumulant expansion at second order, which (as we will see) yields $1/z$ corrections in the effective action \cite{Snoek:2010uq}.
We write the second order approximation of the cumulant expansion as
\begin{equation}
\ln \langle \exp[-i \Delta S + S^{(0)}_\eta] \rangle_{S^{(0)}}
\approx
-i S^{(0)}_{\textrm{eff}} + S^{(0)}_{\textrm{eff}, \eta}
\, ,
\end{equation}
collecting the source-free terms in the effective action
\begin{multline}
S^{(0)}_{\textrm{eff}} =
\int_\mathcal{C} \! dt \, \langle \Delta H(t) \rangle^{(c)}_{S^{(0)}}
\\
+ \frac{i}{2} \iint_\mathcal{C} \! dt \, dt' \,
\langle \Delta H(t) \Delta H(t') \rangle^{(c)}_{S^{(0)}}
\, , \label{eq:S0effCumulant}
\end{multline}
and the terms containing source fields $\eta$ in the effective \emph{source} action
\begin{multline}
S^{(0)}_{\textrm{eff}, \eta} =
\int_\mathcal{C} dt \, \langle H^{(0)}_\eta(t) \rangle^{(c)}_{S^{(0)}}
- \frac{i}{2} \iint_\mathcal{C} dt \, dt' \,
\\
\times \Big[
\langle \Delta H(t) H^{(0)}_\eta (t') \rangle^{(c)}_{S^{(0)}}
+ \langle H^{(0)}_\eta (t) \Delta H(t') \rangle^{(c)}_{S^{(0)}}
\Big]
\\
+ \frac{1}{2} \iint_\mathcal{C} dt \, dt' \,
\langle H^{(0)}_\eta(t) H^{(0)}_\eta(t') \rangle^{(c)}_{S^{(0)}}
\, . \label{eq:S0effetaCumulant}
\end{multline}
Hence, by truncating the expansion at second order we obtain an effective action $S_{\textrm{eff}}$ and generating functional $\mathcal{Z}_{\textrm{eff}}[\eta]$ according to
\begin{multline}
\frac{\mathcal{Z}[\eta]}{\mathcal{Z}^{(0)}} =
\textrm{Tr}_0 \left[ \mathcal{T}_\C
e^{-iS_0 + S_{0, \eta}}
\langle e^{-i \Delta S + S^{(0)}_\eta} \rangle_{S^{(0)}}
\right]
\\ \approx
\textrm{Tr}_0 \left[ \mathcal{T}_\C
\exp[-iS_0 - iS^{(0)}_{\textrm{eff}} + S_{0, \eta} + S^{(0)}_{\textrm{eff}, \eta}]
\right]
\\
\equiv
\textrm{Tr}_0 \left[ \mathcal{T}_\C
\exp[-iS_{\textrm{eff}}] \right]
= \mathcal{Z}_{\textrm{eff}}[\eta] \, .
\end{multline}
\subsection{Explicit 2nd order form}
To obtain the explicit form for the local effective action
\begin{equation}
S_{\textrm{eff}} = S_0 + S^{(0)}_{\textrm{eff}} + i S_{0,\eta} + i S^{(0)}_{\textrm{eff},\eta}
\label{eq:SeffComponents}
\end{equation}
we need to look into the details of the expansion giving the actions $S^{(0)}_{\textrm{eff}}$ and $S^{(0)}_{\textrm{eff},\eta}$. At a later stage, we will also make use of the effective generating functional in order to arrive at the contour generalization of the (self-consistent) B-DMFT effective action, previously derived for equilibrium in \cite{Anders:2011uq, Anders:2011fk}.
The operators appearing in the expansion of $S^{(0)}_{\textrm{eff}}$ and $S^{(0)}_{\textrm{eff},\eta}$ [Eqs.\ (\ref{eq:S0effCumulant}) and (\ref{eq:S0effetaCumulant})] are
\begin{align}
\Delta H &
= -J\sum_{\langle 0, i \rangle} \mbf{b}^{\dagger}_i \mbf{b}^{\phantom{\dagger}}_0
= -J\sum_{\langle 0, i \rangle} \mbf{b}^{\dagger}_0 \mbf{b}^{\phantom{\dagger}}_i \, , \label{eq:DeltaH}\\
H^{(0)}_\eta &
= \sum_{i \ne 0} \mbf{b}^{\dagger}_i \boldsymbol{\eta}^{\phantom{\dagger}}_i
= \sum_{i \ne 0} \boldsymbol{\eta}^{\dagger}_i \mbf{b}^{\phantom{\dagger}}_i \, , \label{eq:H0eta}
\end{align}
where in the last steps we have used the hermitian property of Nambu creation-annihilation operator products [Eq.\ (\ref{eq:NambuCommutation})].
Hence the first order expectation values take the form
\begin{align}
\langle \Delta H(t) \rangle^{(c)}_{S^{(0)}} &
= -J\sum_{\langle 0, i \rangle} \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}} \mbf{b}^{\phantom{\dagger}}_0(t) \, ,\\
\langle H^{(0)}_\eta(t) \rangle^{(c)}_{S^{(0)}} &
= \sum_{i \ne 0} \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}} \boldsymbol{\eta}^{\phantom{\dagger}}_i(t) \, .
\end{align}
The second order terms can be obtained using the two different ways of expressing the operators in Eqs.\ (\ref{eq:DeltaH}) and (\ref{eq:H0eta}) in order to arrive at Nambu response function expressions as in Eq.\ (\ref{eq:NambuGf}).
The second order term of $S^{(0)}_{\textrm{eff}}$ in Eq.\ (\ref{eq:S0effCumulant}) reads
\begin{multline}
\langle \Delta H(t) \Delta H(t') \rangle^{(c)}_{S^{(0)}}
\\ =
\mbf{b}^{\dagger}_0(t)
\Big[ \sum_{ \langle 0,i \rangle, \, \langle 0, j \rangle }
J \langle \mbf{b}^{\phantom{\dagger}}_i(t) \mbf{b}^{\dagger}_j(t') \rangle^{(c)}_{S^{(0)}} J
\Big] \mbf{b}^{\phantom{\dagger}}_0(t')
\\ =
i \mbf{b}^{\dagger}_0(t)
\Big[ \sum_{ \langle 0,i \rangle, \, \langle 0, j \rangle }
J \mbf{G}^{(0)}_{ij}(t, t') J
\Big] \mbf{b}^{\phantom{\dagger}}_0(t')
\\ =
i \mbf{b}^{\dagger}_0(t)
\boldsymbol{\Delta}(t, t') \mbf{b}^{\phantom{\dagger}}_0(t')
\, ,
\end{multline}
where we have introduced the connected single particle Green's function $\mbf{G}^{(0)}_{ij}(t, t') = -i \langle \mbf{b}^{\phantom{\dagger}}_i(t) \mbf{b}^{\dagger}_j(t') \rangle^{(c)}_{S^{(0)}}$ of the lattice with cavity and the total hybridization function $\boldsymbol{\Delta}$ of the zeroth lattice site
\begin{equation}
\boldsymbol{\Delta}(t, t') =
\sum_{ \langle 0,i \rangle, \, \langle 0, j \rangle }
J \mbf{G}^{(0)}_{ij}(t, t') J
\, . \label{eq:bDeltaDefInG0}
\end{equation}
Hence, the source-free action $S^{(0)}_{\textrm{eff}}$ can be written as
\begin{multline}
S^{(0)}_{\textrm{eff}} =
-J \int_\mathcal{C} dt \sum_{\langle 0,i \rangle} \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}} \mbf{b}^{\phantom{\dagger}}_0(t)
\\
+ \frac{1}{2} \iint_\mathcal{C} dt \, dt' \,
\mbf{b}^{\dagger}_0(t) \boldsymbol{\Delta}(t, t') \mbf{b}^{\phantom{\dagger}}_0(t')
\, . \label{eq:S0eff}
\end{multline}
\subsubsection{Local effective source action}
Next we consider the second order terms of the effective source action $S^{(0)}_{\textrm{eff},\eta}$ [Eq.\ (\ref{eq:S0effetaCumulant})]. In terms of $\mbf{G}^{(0)}_{ij}(t, t')$ the quadratic source term reads
\begin{multline}
\langle H^{(0)}_\eta(t) H^{(0)}_\eta(t') \rangle^{(c)}_{S^{(0)}}
\\ =
i \sum_{i \ne 0 , \, j \ne 0}
\boldsymbol{\eta}^{\dagger}_i(t) \mbf{G}^{(0)}_{ij}(t, t') \boldsymbol{\eta}^{\phantom{\dagger}}_j(t') \, ,
\end{multline}
and the first mixed term becomes
\begin{multline}
\langle \Delta H(t) H^{(0)}_\eta(t') \rangle^{(c)}_{S^{(0)}}
\\ =
-i \mbf{b}^{\dagger}_0(t)
\sum_{\langle 0, i \rangle , \, j \ne 0} J \mbf{G}^{(0)}_{ij}(t, t') \boldsymbol{\eta}^{\phantom{\dagger}}_j(t')
\, .
\end{multline}
By an interchange of integration variables it is possible to show that the other mixed term gives an equal contribution.
Collecting all the terms we arrive at the final expression for the local effective source action
\begin{multline}
S^{(0)}_{\textrm{eff}, \eta} =
\int_\mathcal{C} dt \sum_{i \ne 0} \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}} \boldsymbol{\eta}^{\phantom{\dagger}}_i(t)
\\
+ \frac{i}{2} \iint_\mathcal{C} dt \, dt' \,
\sum_{i \ne 0 , \, j \ne 0}
\boldsymbol{\eta}^{\dagger}_i(t) \mbf{G}^{(0)}_{ij}(t, t') \boldsymbol{\eta}^{\phantom{\dagger}}_j(t')
\\
- \iint_\mathcal{C} dt \, dt' \,
\mbf{b}^{\dagger}_0(t)
\sum_{\langle 0, i \rangle , \, j \ne 0} J \mbf{G}^{(0)}_{ij}(t, t') \boldsymbol{\eta}^{\phantom{\dagger}}_j(t')
\, ,
\label{eq:S0effEta}
\end{multline}
\subsection{Local anomalous term}
We see in Eq.\ (\ref{eq:S0eff}) that the symmetry breaking of the infinite lattice system induces a local symmetry breaking term on the zeroth lattice site. The strength of the symmetry breaking field is however determined by the anomalous expectation values $\langle \mbf{b}^{\dagger}_i(z) \rangle^{(c)}_{S^{(0)}}$ on all sites $i$ \emph{neighboring the cavity}.
For finite coordination numbers $z$ the removal of the cavity site affects the neighboring sites, hence this expectation value is \emph{not} equal to that of the original homogeneous system \cite{Snoek:2010uq}
\begin{equation}
\langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}}
\ne \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S}
\approx \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S_{\textrm{eff}}} \, .
\end{equation}
To determine the difference between $\langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}}$ and $\langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S_{\textrm{eff}}}$ we calculate the latter using the effective generating functional $\mathcal{Z}_{\textrm{eff}}$ and the Nambu generalization of Eq.\ (\ref{eq:ZResponseFunction})
\begin{multline}
\langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S_{\textrm{eff}}}
= \frac{\partial}{\partial \boldsymbol{\eta}^{\phantom{\dagger}}_i(t)} \ln \mathcal{Z}_{\textrm{eff}}[\eta] |_{\eta = 0}
= \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}}
\\ - \int_\mathcal{C} dt' \,
\langle \mbf{b}^{\dagger}_0(t') \rangle^{(c)}_{S_{\textrm{eff}}}
\sum_{\langle 0, j \rangle} J \mbf{G}^{(0)}_{ji}(t', t)
\, .
\end{multline}
Thus the local anomalous term of $S^{(0)}_{\textrm{eff}}$ in Eq.\ (\ref{eq:S0eff}) can be rewritten, using only expectation values with respect to $S_{\textrm{eff}}$, as
\begin{multline}
-J \int_\mathcal{C} dt \sum_{\langle 0,i \rangle} \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S^{(0)}} \mbf{b}^{\phantom{\dagger}}_0(t)
=
-J \int_\mathcal{C} dt \sum_{\langle 0,i \rangle}
\Big[
\langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S_{\textrm{eff}}}
\\ +
\int_\mathcal{C} dt' \,
\langle \mbf{b}^{\dagger}_0(t') \rangle^{(c)}_{S_{\textrm{eff}}}
\sum_{\langle 0, j \rangle} J \mbf{G}^{(0)}_{ji}(t', t)
\Big]
\mbf{b}^{\phantom{\dagger}}_0(t)
\\ \shoveleft =
\int_\mathcal{C} dt
\Big[
-zJ \boldsymbol\Phi^\dagger_0(t)
- \int_\mathcal{C} dt' \,
\boldsymbol\Phi^\dagger_0(t')\boldsymbol{\Delta}(t', t)
\Big] \mbf{b}^{\phantom{\dagger}}_0(t) \, ,
\label{eq:EffLocalAnomalousTerm}
\end{multline}
where in the last step, we have introduced the local anomalous amplitude $\boldsymbol\Phi_0^\dagger(t) = \langle \mbf{b}^{\dagger}_0(t) \rangle^{(c)}_{S_{\textrm{eff}}}$ and assumed translational invariance $\langle \mbf{b}^{\dagger}_0(t) \rangle^{(c)}_{S_{\textrm{eff}}} = \langle \mbf{b}^{\dagger}_i(t) \rangle^{(c)}_{S_{\textrm{eff}}}$.
\subsection{Local effective action}
Substituting Eqs.\ (\ref{eq:S0eff}) and (\ref{eq:S0effEta}) into Eq.\ (\ref{eq:SeffComponents}), rewriting the local symmetry breaking using Eq.\ (\ref{eq:EffLocalAnomalousTerm}) and setting the sources to zero $\eta = 0$, we obtain the BDMFT local effective action for the Bose-Hubbard model
\begin{multline}
S_{\textrm{eff}} =
\int_\mathcal{C} dt \left(
-\mu n(t)
+ \frac{U}{2} n(t) ( n(t) - 1 )
\right)
\\ +
\int_\mathcal{C} dt
\Big[
-zJ \boldsymbol\Phi^\dagger(t)
- \int_\mathcal{C} dt' \,
\boldsymbol\Phi^\dagger(t')\boldsymbol{\Delta}(t', t)
\Big] \mbf{b}(t)
\\ + \frac{1}{2} \iint_\mathcal{C} dt \, dt' \,
\mbf{b}^{\dagger}(t) \boldsymbol{\Delta}(t, t') \mbf{b}(t') \, ,
\label{eq:EffectiveActionWithPhiEff}
\end{multline}
where we have dropped site indices and the complex field $\boldsymbol\Phi^\dagger(t)$ is self-consistently defined as $\boldsymbol\Phi^\dagger(t) = \langle \mbf{b}^{\dagger}(t) \rangle_{S_{\textrm{eff}}}$.
\subsubsection{Equilibrium form}
On
the imaginary time branch the field is constant $\boldsymbol\Phi^\dagger(\tau) = \boldsymbol\Phi^\dagger$, the hybridization function is time translational invariant $\boldsymbol{\Delta}(\tau, \tau') = \boldsymbol{\Delta}(\tau - \tau')$, and the action simplifies to
\begin{multline}
S_{\textrm{eff}} =
\int_0^\beta d\tau \left(
- \mu n(\tau)
+ \frac{U}{2} n(\tau) ( n(\tau) - 1 )
\right)
\\ +
\boldsymbol\Phi^\dagger
\Big[
-zJ
-\int_0^\beta d\bar{\tau} \,
\boldsymbol{\Delta}(\bar{\tau})
\Big]
\int_0^\beta d\tau
\mbf{b}(\tau)
\\ + \frac{1}{2} \iint_0^\beta d\tau \, d\tau' \,
\mbf{b}^{\dagger}(\tau) \boldsymbol{\Delta}(\tau - \tau') \mbf{b}(\tau') \, ,
\end{multline}
in agreement with Refs.\ \onlinecite{Anders:2011uq, Anders:2011fk}, up to a minus sign on the hybridization function due to a different notation.
\subsection{One-loop correction in $1/z$}
To see that the BDMFT effective action in Eq.\ (\ref{eq:EffectiveActionWithPhiEff}) is a one-loop correction in the inverse coordination number $1/z$ one must study the scaling of its terms. %
The only non-trivial contribution comes from the hybridization function $\boldsymbol{\Delta} = J^2 \sum_{ \langle 0,i \rangle, \, \langle 0, j \rangle } \mbf{G}^{(0)}_{ij}$ [Eq.\ (\ref{eq:bDeltaDefInG0})].
On a graph without loops, such as the Bethe graph, the sum over nearest neighbors contains no cross-terms $\mbf{G}^{(0)}_{ij} = \delta_{ij} \mbf{G}^{(0)}_{ii}$, and the sum simplifies to
\begin{equation}
\boldsymbol{\Delta}(t, t') = z J^2 \mbf{G}^{(0)}_{ii}(t, t') \, .
\end{equation}
On more general lattices, the power counting in $z$ gives the same leading order result, but is more elaborate \cite{Georges:1996aa}.
Substituting this result into Eq.\ (\ref{eq:EffectiveActionWithPhiEff}) and making a $J \rightarrow J/z$ rescaling of the hopping gives the rescaled action
\begin{multline}
\tilde{S}_{\textrm{eff}} =
\int_\mathcal{C} dt \left(
-\mu n(t)
+ \frac{U}{2} n(t) ( n(t) - 1 )
\right)
\\ +
\int_\mathcal{C} dt
\Big[
-J \boldsymbol\Phi^\dagger(t)
- \frac{J^2}{z} \int_\mathcal{C} dt' \,
\boldsymbol\Phi^\dagger(t') \mbf{G}^{(0)}_{ii}(t', t)
\Big] \mbf{b}(t)
\\ + \frac{J^2}{2z} \iint_\mathcal{C} dt \, dt' \,
\mbf{b}^{\dagger}(t) \mbf{G}^{(0)}_{ii}(t, t') \mbf{b}(t') \, ,
\end{multline}
which makes it evident that the terms containing $\mbf{G}^{(0)}_{ii}$, i.e.\ the hybridization terms in the BDMFT effective action [Eq.\ (\ref{eq:EffectiveActionWithPhiEff})] corresponds to a $1/z$ correction of the mean-field action (obtained by setting $\boldsymbol{\Delta} = 0$).
\subsection{Second order fluctuation expansion}
While BDMFT can bee seen as a one-loop expansion in the inverse coordination number it is also a second order expansion in the condensate fluctuations, as discussed in Ref.\ \onlinecite{Anders:2011uq}. This can be made explicit by rewriting the terms containing the hybridization in the effective action using the fluctuation operators $\delta\mbf{b}$, defined as $\delta\mbf{b} \equiv \mbf{b} - \mbf{\Phi}$ with $\langle \delta\mbf{b} \rangle = \mbf{0}$. Inserting these in Eq.~(\ref{eq:EffectiveActionWithPhiEff}) yields
\begin{multline}
S_{\textrm{eff}} =
\int_\mathcal{C} dt \left(
-\mu n(t)
+ \frac{U}{2} n(t) ( n(t) - 1 ) - zJ \mbf{\Phi}^\dagger(t) \mbf{b}(t)
\right)
\\
+ \frac{1}{2} \iint_\mathcal{C} dt \, dt' \,
\delta\mbf{b}^\dagger(t) \boldsymbol{\Delta}(t, t') \delta\mbf{b}(t') \, ,
\end{multline}
where the hybridization term is the exact 2nd order contribution of the fluctuations.
Hence BDMFT correctly describes the deep superfluid where fluctuations are suppressed, i.e.\ the weakly interacting Bose gas limit (WIBG) \cite{Capogrosso-Sansone:2010vn}.
\section{Introduction}
Cold atomic gases trapped in an optical lattice provide a unique play-ground to explore equilibrium and nonequilibrium properties of interacting many-particle systems \cite{Morsch:2006vn, Bloch:2008uq}. They enable an almost ideal realization of the low-energy effective Hamiltonians (the fermionic and bosonic Hubbard models \cite{Hubbard:1963aa, Fisher:1989kl}) which have been studied in the condensed matter context for a long time, and whose properties are still not yet fully understood. A big advantage of cold atoms, as compared to condensed matter systems, is that interaction parameters can be tuned almost arbitrarily, and that the lattice spacings and characteristic time-scales are much larger \cite{Jaksch:1998vn}. For bosonic atoms, the Mott insulating and superfluid regime can easily be accessed \cite{Morsch:2006vn} and the experimental control is so precise that the use of cold atoms as ``quantum simulators" becomes a realistic option \cite{Trotzky:2010fk} (for a recent review see Ref.\ \onlinecite{Kennett:2013fk}).
A particularly interesting aspect of cold atom experi\-ments is the possibility to study the time-evolution of interacting many-body systems \cite{Greiner:2002fk,Sebby-Strabley:2007vn, Will:2010uq, Bakr:2010dq, Bissbort:2011bs, Endres:2012bs, Cheneau:2012ys, Trotzky:2012kx, Braun:2014kx}. This was beautifully demonstrated in the seminal work by Greiner {\it et al.~}\cite{Greiner:2002fk}, who measured the condensate collapse-and-revival oscillations after a quench in a Bose-Hubbard system from the superfluid to the Mott regime. In contrast to equilibrium, where the phase diagram and correlation functions of the Bose-Hubbard model \cite{Capogrosso-Sansone:2007lh} can be computed accurately using Monte Carlo simulations \cite{Pollet:2012ly}, the real-time evolution of interacting bosonic lattice systems is a big computational challenge.
In one dimension, density matrix renormalization group (DMRG) methods \cite{Schollwock:2005ly} can be used to simulate the time-evolution after a quench on relatively large lattices, but a rapid entanglement growth limits the accessible time-scale \cite{Trotzky:2012kx}. Still, DMRG calculations have provided important insights into the short time dynamics, as measured in 1D optical lattices \cite{Cheneau:2012ys, Trotzky:2012kx, Braun:2014kx}.
Kollath {\it et al.~}\cite{Kollath:2007ys} used non-local correlators to study relaxation and thermalization. They showed that an initially superfluid system is trapped in a nonthermal steady state after quenching the interaction deep into the Mott regime, while thermalization occurs after quenches to intermediate interactions.
Also the eigenstate thermalization hypothesis has been explored \cite{Roux:2009ys, Sorg:2014vn} and debated \cite{Rigol:2010zr, Roux:2010ly} in this context.
A more recent development is the time-dependent variational Monte Carlo (tVMC) approach that shows good agreement with DMRG in 1D without being limited in time \cite{Carleo:2012zr}. It has also been applied to 2D systems and is not inherently limited to any dimensionality \cite{Carleo:2014ly}. While tVMC is well suited for studying the spread of correlations, it is a method that treats finite systems,
which complicates the study of thermalization \cite{Biroli:2010qf}.
In three dimensions, perturbation theory \cite{Trefzger:2011uq, Kennett:2011fk, Dutta:2012kx, Dutta:2014uq}, and Gutzwiller mean-field (MF) \cite{Huber:2007ys, Wolf:2010bh, Sciolla:2010uq, Sciolla:2011kx, Snoek:2011hc, Krutitsky:2011qf} calculations have been performed. Both work in specific regions of the phase diagram, but generally fail to describe finite temperature relaxation and thermalization phenomena.
Hence, while being accessible experimentally \cite{Braun:2014kx}, out-of-equilibrium phenomena in the three dimensional Bose-Hubbard model remain largely unexplored \cite{Trotzky:2012kx, Cheneau:2012ys, Braun:2014kx} from the theoretical point of view.
Describing the generic relaxation phenomena and nonthermal transient states, as well as mapping out the different dynamical regimes of this model is fundamental to our understanding of nonequilibrium lattice bosons.
A clear picture of the nonequilibrium properties of the homogeneous bulk-system is also important for the interpretation of more complicated experimental set-ups.
For example, one open question is whether damped superfluid collapse-and-revival oscillations are a dynamical feature of the homogeneous system, or an effect of the trapping potential or other processes not considered in the Bose-Hubbard description \cite{Greiner:2002fk, Will:2010uq}.
A computationally tractable and promising scheme, which allows to address such issues, is the nonequilibrium generalization of bosonic dynamical mean field theory (BDMFT).
This method is formulated in the thermodynamical limit, and thus enables the study of relaxation and thermalization phenomena in infinite systems \cite{Aoki:2014kx}.
The equilibrium version of BDMFT \cite{Byczuk:2008nx, Hubener:2009cr, Anders:2010uq, Anders:2011uq} produces phase diagrams, condensate fractions, and correlation functions with remarkable accuracy \cite{Anders:2011uq}.
While the extension of this formalism to nonequilibrium situations is analogous to the fermionic case \cite{Aoki:2014kx}, and essentially involves the replacement of the imaginary-time interval by a Kadanoff-Baym contour, there are a number of practical challenges.
The most important one is the development of a suitable bosonic impurity solver. The exact continuous-time quantum Monte Carlo (CT-QMC) impurity solver of Ref.~\onlinecite{Anders:2010uq} cannot easily be applied to nonequilibrium problems, because of a dynamical sign problem \cite{Werner:2009tg}, while exact diagonalization based solvers are even more limited than in the fermionic case \cite{Gramsch:2013fk}, due to the larger local Hilbert space. Weak-coupling perturbation theory is not an option if one is interested in Mott physics. Instead, we will develop and benchmark an impurity solver based on the lowest order strong-coupling perturbation theory, i.e.\ the non-crossing approximation (NCA) \cite{Keiter:1971hc}.
As a first application of this new scheme, we will map out the different dynamical regimes of both the symmetric and symmetry broken states, searching for thermalization and trapping phenomena after a quench of the interaction parameter.
This paper is organized as follows:
In section \ref{sec:Theory} we give an overview of the Bose-Hubbard model, the nonequilibrium generalization of BDMFT [Sec.\ \ref{sec:BDMFT}], the NCA impurity solver [Sec.\ \ref{sec:NCA}], the energy calculations [Sec.\ \ref{sec:Energy}], and our numerical implementation [Sec.\ \ref{sec:Numerics}]. In Sec.\ \ref{sec:Results} we first present benchmark calculations showing density and energy conservation and discuss the lowest order spectral moments [Sec.\ \ref{sec:Benchmark}]. The dynamical regimes in the normal phase are mapped out in Sec.\ \ref{sec:Mott}. In Sec.\ \ref{sec:SuperFluid} we consider superfluid initial states, and after an overview of the relaxation regimes in Sec.\ \ref{sec:SuperFluidRegimes}, we study the dynamics for short times in Sec.\ \ref{sec:ShortTimeDynamics}, and long times in Sec.\ \ref{sec:LongTimeDynamics}. The findings are summarized in Sec.\ \ref{sec:SuperFluidPhaseDiagram} in the form of a nonequilibrium ``phase diagram''. Sec.\ \ref{sec:Conclusions} is devoted to conclusions. We also provide a derivation of nonequilibrium BDMFT in Appendix \ref{app:BDMFT}, and discuss the details of the Nambu generalization of NCA in Appendix \ref{app:NCA}.
\section{Theory} \label{sec:Theory}
We consider the simplest model for bosonic atoms in an optical lattice, namely the Bose-Hubbard model \cite{Fisher:1989kl, Jaksch:1998vn}
\begin{equation}
H = -J \sum_{\langle i,j \rangle} (b^\dagger_i b_j + b^\dagger_j b_i)
+\frac{U}{2} \sum_i \hat{n}_i(\hat{n}_i-1) - \mu \sum_i \hat{n}_i
\, , \label{eq:Model}
\end{equation}
where $b^\dagger_i$ ($b_i$) and $\hat{n}_i$ are the bosonic creation (annihilation) and number operators acting on site $i$, $\mu$ is the chemical potential, and $U$ the local pair interaction which competes with the nearest neighbor hopping $J$ that we take as our unit of energy.
\subsection{Nonequilibrium bosonic dynamical mean-field theory} \label{sec:BDMFT}
By extending the equilibrium bosonic dynamical mean field theory (BDMFT) \cite{Anders:2010uq, Anders:2011uq} to the three-branch Kadanoff-Baym contour $\mathcal{C}$ ($0 \! \rightarrow \! t_\text{max} \! \rightarrow \! 0 \! \rightarrow \! -i\beta$) \cite{Aoki:2014kx, Stefanucci:2013oq}, we obtain the bosonic impurity action
\begin{align}
\mathcal{S}_{\textrm{imp}} & =
\int_\mathcal{C} dt\,\Big( -\mu(t)\hat{n}(t) + \frac{U}{2}\hat{n}(t)(\hat{n}(t)-1) \Big)
\label{eq:Seff}\\
&-\int_\mathcal{C} dt \, \mbf{\Phi}^\dagger_{\textrm{eff}}(t)\mbf{b}(t)
+ \frac{1}{2} \iint_\mathcal{C} dt\,dt'\, \mbf{b}^\dagger(t)\mathcal{D}(t, t')\mbf{b}(t')
\, , \nonumber
\end{align}
where $\mbf{b}^\dagger$ is the Nambu spinor $\mbf{b}^\dagger = (b^\dagger, b)$, $\mathcal{D}(t, t')$ the hybridization function, and $\mbf{\Phi}^\dagger_{\textrm{eff}}$ the effective symmetry breaking field,
which is defined in terms of $\mathcal{D}$, the local condensate fraction $\mbf{\Phi}^\dagger = (\phi^*, \phi)$, and the lattice coordination number $z$ as
\begin{equation}
\mbf{\Phi}^\dagger_{\textrm{eff}}(t) = zJ\mbf{\Phi}^\dagger(t) + \int_\mathcal{C} dt'\, \mbf{\Phi}^\dagger(t')\mathcal{D}(t', t) \, . \label{phieff}
\end{equation}
For a detailed derivation of this action, see App.\ \ref{app:BDMFT}. Note that the single-particle fluctuations (the $\mathcal{D}$ term in Eq.\ (\ref{eq:Seff}) and Eq.\ (\ref{phieff})) enter as a correction to the mean-field action \cite{Sachdev:1999fk}, which would be obtained by taking the infinite dimensional limit $z \rightarrow \infty$ at fixed $zJ$ (or analogously $\mathcal{D} \rightarrow 0$) \cite{Anders:2011uq}.
The solution of the impurity model yields the connected impurity Green's function
\begin{equation*}
\mbf{G}(t,t') =
-i\langle \mathcal{T}_\mathcal{C} \mbf{b}(t) \mbf{b}^\dagger(t')\rangle + i\mbf{\Phi}(t)\mbf{\Phi}^\dagger(t') \, ,
\end{equation*}
where $\mathcal{T}_\mathcal{C}$ is the time-ordering operator on the contour $\mathcal{C}$, and the local condensate fraction is
\begin{equation*}
\mbf{\Phi}(t)=\langle \mbf{b}(t) \rangle \, .
\end{equation*}
The BDMFT self-consistency loop is closed by computing the lattice Green's function $\mbf{G}_{\mbf{k}}$ from $\mbf{G}$, and then expressing the hybridization function $\mathcal{D}$ in terms of the local lattice Green's function $\mbf{G}_{L} = \frac{1}{N_{\mbf{k}}}\sum_{\mbf{k}} \mbf{G}_{\mbf{k}}$ (at self consistency $\mbf{G}_{L} = \mbf{G}$) \cite{Aoki:2014kx}.
In the present study we employ the simplified self-consistency relation
\begin{equation*}
\mathcal{D}(t, t') = (3J)^2 \mbf{G}(t, t') \, ,
\end{equation*}
and set $z=6$, which corresponds to a non-interacting semi-circular density of states (DOS) with the same bandwidth $W = 12J$ and lattice coordination number $z$ as the 3D cubic lattice with nearest neighbor hopping $J$.
\subsection{Non-crossing approximation impurity solver}
\label{sec:NCA}
The previous BDMFT equilibrium studies employed a hybridization expansion CT-QMC impurity solver \cite{Anders:2010uq, Anders:2011uq}. However, the extension of this technique to the contour-action in Eq.\ (\ref{eq:Seff}) does not look promising, because the dynamical sign problem from the expansion along the real-time branches \cite{Werner:2010pb} will add to the inherent sign problem of the hybridization expansion (in the superfluid regime).
We therefore solve the BDMFT effective impurity action using the first order self-consistent strong coupling expansion.
The generalization of strong coupling expansions to real-time impurity problems has been presented in Ref.~\cite{Eckstein:2010fk}. To treat the BDMFT effective action in Eq.\ (\ref{eq:Seff}) we have generalized the formalism to systems with symmetry breaking, as discussed in App.~\ref{app:NCA}.
In short we follow the standard procedure and introduce pseudo-particle second quantization operators $p^{\phantom{\dagger}}_\Gamma$ and $p^\dagger_\Gamma$ for each local occupation number many-body state $| \Gamma \rangle$.
This maps the local Hamiltonian to a quadratic term $\sum_{\Gamma\Gamma'} \hat{H}(t)_{\Gamma\Gamma'} p^\dagger_\Gamma p^{\phantom{\dagger}}_{\Gamma'}$, while the hybridization $\boldsymbol{\Delta}$ turns into a pseudo-particle interaction.
Expanding to first order in $\boldsymbol{\Delta}$ gives the NCA of Ref.\ \onlinecite{Eckstein:2010fk} generalized to Nambu formalism.
\begin{figure}
\begin{flushleft}
\includegraphics[scale=1] {figure1a.pdf} \\
\includegraphics[scale=1] {figure1b.pdf}
\end{flushleft}
\caption{\label{fig:NCA}
NCA diagram representations of a) the pseudo-particle self-energy $\hat{\Sigma}$, and b) the single-particle Green's function $\mbf{G}_{\gamma \nu}$.
} \end{figure}
The corresponding NCA pseudo-particle self-energy $\hat{\Sigma} = \hat{\Sigma}_{\Gamma\Gamma'}$ consists of the two shell diagrams with a directed hybridization line (see Fig.~\ref{fig:NCA} and App.\ \ref{app:NCASigma})
\begin{align}
\hat{\Sigma}(t, t') =
\frac{i}{2} \sum_{\gamma \nu} \Big( &
\mathcal{D}_{\gamma \nu}(t, t') \left[ \mbf{b}^\dagger_\gamma \hat{G}(t, t') \mbf{b}_\nu \right] + \nonumber \\ &
\mathcal{D}_{\nu \gamma}(t', t) \left[ \mbf{b}_\gamma \hat{G}(t, t') \mbf{b}^\dagger_\nu \right] \Big) \, ,
\label{eq:NCASigma}
\end{align}
where $\hat{G} = \hat{G}_{\Gamma\Gamma'}$ is the pseudo-particle Green's function, $\gamma$ and $\nu$ are Nambu indices, and $\mbf{b}_\gamma$ is the tensor $(\mbf{b}_\gamma)_{\Gamma\Gamma'} = \langle \Gamma | \mbf{b}_\gamma | \Gamma' \rangle$ (operator products are implicit matrix products).
The pseudo-particle Dyson equation takes the form
\begin{equation*}
( i \partial_t + \hat{H}(t) )\hat{G} - \mbox{$\hat{\Sigma} \mathrlap{\kern0.085em{\circlearrowright}}\ast \hat{G}$} = 0,
\end{equation*}
where $\hat{H}(t)$ is the static part in Eq.\ (\ref{eq:Seff}), $\hat{H}(t) = U(t)(\hat{n}^2 - \hat{n})/2 - \mu(t)\hat{n} - \mbf{\Phi}^\dagger_{\textrm{eff}}(t) \mbf{b}$, and $\mbox{$\hat{\Sigma} \mathrlap{\kern0.085em{\circlearrowright}}\ast \hat{G}$}$ denotes cyclic convolution on $\mathcal{C}$, $(\mbox{$\hat{\Sigma} \mathrlap{\kern0.085em{\circlearrowright}}\ast \hat{G}$})(t, t') = \int_{t' \prec \bar{t} \prec t} d\bar{t} \, \hat{\Sigma}(t, \bar{t}) \, \hat{G}(\bar{t}, t')$ \cite{Eckstein:2010fk}.
Within NCA, $\hat{G}$ and $\hat{\Sigma}$ are calculated self-consistently, and local observables are determined from the reduced local density matrix $\hat{\rho}(t) = i\hat{G}^<(t,t)$, yielding the local condensate as
\begin{equation*}
\mbf{\Phi}_\gamma(t) =
\langle \mbf{b}_\gamma(t) \rangle = \textrm{Tr}_\Gamma [ \mbf{b}_\gamma \hat{\rho}(t) ],
\end{equation*}
while response functions must be determined diagrammatically. In particular, the connected single-particle impurity Green's function $\mbf{G}$ is obtained from the bubble diagram without hybridization insertions
(see Fig.\ \ref{fig:NCA} and App.\ \ref{app:NCAGsp})
\begin{multline}
\mbf{G}_{\gamma \nu}(t, t') = \\
i \textrm{Tr}_\Gamma \left[ \hat{G}(t', t) \mbf{b}_\gamma \, \hat{G}(t, t') \mbf{b}^\dagger_\nu \right]
+ i \mbf{\Phi}_\gamma(t) \mbf{\Phi}^\dagger_\nu(t') .
\label{eq:NCAGsp}
\end{multline}
\subsection{Total energy components} \label{sec:Energy}
The total energy $E_t$ of the system, is the sum of the (connected) kinetic energy $E_k$, the condensate energy (or disconnected kinetic energy) $E_c$, and the local interaction energy $E_i$, $E_t = E_k + E_c + E_i$.
Using $\mbf{G}$ and $\mathcal{D}$, $E_k$ is given by \cite{Aoki:2014kx}
\begin{equation*}
E_k(t) = \frac{i}{2} \textrm{Tr} \left[ (\mathcal{D} * \mbf{G})^<(t, t) \right],
\end{equation*}
$E_c$ depends on $\phi(t) = \langle b(t) \rangle = \textrm{Tr}_\Gamma[ b \, \hat{\rho}(t) ]$ as
\begin{equation*}
E_c(t) = -zJ(t) | \phi(t) |^2,
\end{equation*}
and $E_i$ can be written in terms of $\langle \hat{n}^2 \rangle(t) = \textrm{Tr}_\Gamma[ \hat{n}^2 \, \hat{\rho}(t) ]$ and $\langle \hat{n} \rangle(t) = \textrm{Tr}_\Gamma[ \hat{n} \, \hat{\rho}(t) ]$ as
\begin{equation*}
E_i(t) = U(t) ( \langle \hat{n}^2 \rangle(t) - \langle \hat{n} \rangle(t) )/2.
\end{equation*}
\subsection{Numerical implementation} \label{sec:Numerics}
We solve the pseudo-particle Dyson equation using a fifth order multi-step method \cite{Brunner:1986ff, Eckstein:2010fk} on an uniformly discretized time grid.
To ensure negligible real-time discretization errors we monitor the total energy and density, which both are constants of motion of the conserving NCA \cite{Eckstein:2010fk} (the gauge property $\mu \rightarrow \mu + \delta\mu(t) \Rightarrow b \rightarrow b e^{-i \int_0^t d\bar{t} \, \delta\mu(\bar{t})}$ ensures $\partial_t \langle \hat{n} \rangle = 0$).
In principle the local Fock space is unbounded, but for $U>0$ it can safely be truncated, keeping only $N_{\textrm{max}}$ states. The cut-off error is controlled by monitoring the drift in $\textrm{Tr}_\Gamma[ \hat{\rho} ]$ away from unity. Close to the $\langle \hat{n} \rangle = 1$ superfluid transition at $U \gtrsim J$, the results are converged for $N_{\textrm{max}} = 5$ to $11$.
The computational limitations of our real-time BDMFT+NCA implementation are very similar to the real-time fermionic DMFT+NCA case \cite{Eckstein:2010fk}. Memory is \emph{the} limiting factor when working with two time response functions, whose storage size scales quadratically with the number of time steps. The local Fock space in the bosonic case adds one or two orders of magnitude in memory usage, compared to the single-band fermionic case. A further limitation is the quadratic energy dependence of the local occupation number states $| \Gamma \rangle$, scaling with $E_\Gamma \sim \langle \Gamma | U\hat{n}^2 | \Gamma \rangle = Un_\Gamma^2$. This induces a pseudo-particle time dependence $\hat{G}_{\Gamma \Gamma}(t, t') \sim e^{-iUn_\Gamma^2(t-t')}$, which means that including higher occupation number states, by increasing $N_{\textrm{max}}$, also requires a finer time discretization.
\section{Results} \label{sec:Results}
\begin{figure}
\includegraphics[scale=1.0]{figure2.pdf}
\caption{\label{fig:PhaseDiag} (color online) BDMFT superfluid phase boundary for CT-QMC \cite{Anders:2011uq} (blue) and NCA (red) on the 3D cubic lattice, NCA with a semi circular DOS (green), and mean-field theory (black). Panel a) shows the ($J/U$, $\mu/U$) plane at $T=1.5$, and panel b) the ($U$, $T$) plane for $\langle \hat{n} \rangle = 1$.}
\end{figure}
\subsection{Benchmark calculations} \label{sec:Benchmark}
Even though BDMFT neglects spatial fluctuations, the equilibrium results for the 3D Bose-Hubbard model are in good quantitative agreement \cite{Anders:2010uq, Anders:2011uq} with high precision lattice QMC calculations \cite{Capogrosso-Sansone:2007lh} and high-order perturbation theory \cite{Teichmann:2009bh}, for both the phase diagram and local correlation functions.
For example, the critical couplings at the $\langle \hat{n} \rangle = 1$ superfluid-Mott transition are $(J/U)_c = 0.0345 \pm 0.0004$ (BDMFT at $\beta J = 2$), and $(J/U)_c = 0.03408(2)$ (lattice QMC), see Ref.\ \onlinecite{Anders:2011uq} for an explicit comparison of phase diagrams.
To assess the validity of the NCA approximation we compare its superfluid phase boundary for the 3D cubic lattice with the (within BDMFT) exact CT-QMC result, see Fig.\ \ref{fig:PhaseDiag}. It is evident that already this lowest order strong coupling expansion provides a very good approximation with $(J/U)_c \approx 0.0340$ (at $T=1.5$), as expected, considering the success of the linked cluster expansion \cite{Kauch:2012ve}. The simplified self-consistency based on the semi-circular DOS leads to a shift in the phase boundaries (Fig.\ \ref{fig:PhaseDiag}) with $(J/U)_c \approx 0.0378$, but we expect that the qualitative features of the solution, both in and out of equilibrium, remain unchanged.
Note that the Mott phase is only present at integer fillings. Hence, in order to study quenches between the superfluid and Mott insulator, we limit our calculations to $\langle \hat{n} \rangle = 1$. Strictly speaking the Mott insulator exist only at zero temperature, with a smooth crossover to the normal phase, see Fig.\ \ref{fig:PhaseDiag}. However we follow Ref.\ \onlinecite{Capogrosso-Sansone:2007lh} and define the \emph{Mott regime} as the whole region $U > U_c(T=0)$, where the low temperature superfluid phase is absent.
For instantaneous interaction quenches the final total energy $E^{(f)}_{t}$ is given by the initial equilibrium total energy $E^{(i)}_{t}$ and an additional interaction energy contribution $E^{(f)}_{t} = E^{(i)}_{t} + (U_f/U_i - 1) E^{(i)}_{i}$ (due to the sudden change of $U$ from $U_i$ to $U_f$ at $t=0$). Given $E^{(f)}_{t}$ and $U_f$ the effective temperature $T_{\textrm{eff}}$ of the system after thermalization can be determined using separate equilibrium calculations.
The resulting non-equilibrium $(U_{f}, T_{\textrm{eff}})$ pair of a quench can be used to determine the final state after eventual thermalization by direct comparison with the equilibrium $(U, T)$ phase boundaries. This will be used throughout this study in order to produce combined equilibrium $(U, T)$ and non-equilibrium $(U_f, T_{\textrm{eff}})$ ``phase diagrams''.
\begin{figure}
\includegraphics[scale=1]{figure3.pdf}
\caption{\label{fig:SFquenchEnergies} (color online) Time-evolution of energies and observables (lines), and thermal values (thin lines) for the
superfluid to normal phase quench from $U_i=6$ ($T_i=4.5$) to $U_f=21$ ($T_{\textrm{eff}} \approx 9.81$) (left panel),
and time discretization induced drifts $\Delta n = \langle \hat{n}(t) \rangle - \langle \hat{n}(0) \rangle$ and $\Delta E_{t}(t) = E_{t}(t) - E_{t}(0)$ (right panel).} \label{fig_energies}
\end{figure}
\begin{figure*}
\includegraphics[scale=1]
{figure4.pdf}
\caption{\label{fig:UTPhaseDiagram} (color online)
a) Non-equilibrium $(U_f, T_{\textrm{eff}})$ ``phase diagram'' for quenches within the symmetric Mott and normal phases ($|\phi| = 0$) with $\langle \hat{n} \rangle = 1$. While the superfluid state is absent, the equilibrium superfluid $(U, T)$ phase boundary is shown for guidance (dotted line).
The shaded areas indicate the occurrence of rapid thermalization, for the initial interactions $U_i=30$ (cyan area) and $45$ (magenta area) and $T_i \in [3, 18]$. The point of most rapid thermalization is defined as $U_f$ maximizing $|1-\kappa(t=t_{\textrm{max}})|^{-1}$ (diamonds), and the left and right boundaries of the rapid thermalization area correspond to a three-fold decrease from the maxima, as explicitly shown for $U_i=30$ and $T_i=6$ in (b).
For the quenches from $U_i=30$ and $T_i=6$ ending at $(U_f, T_{\textrm{eff}})$ (solid gray line) the real-time evolution of $\kappa(t)$ for $U_f = 4.2$, $10.2$, $15$, and $21$ (blue, green, red, and cyan lines) (see circles in a)), are shown in c) for short times and in d) for long times with $|1-\kappa(t=t_{\textrm{max}})|$ (markers).
} \end{figure*}
BDMFT captures the conversion between interaction, kinetic, and condensate energy, as well as the relaxation to the predicted thermal values (Fig.~\ref{fig_energies}). Despite a nontrivial time-evolution of the individual components the total energy $E_{t}$ and the particle number $\langle \hat{n} \rangle$ is conserved to high accuracy by our 5th order solver (right panel).
We should note, however, that the NCA solution yields an approximate spectral function, as for the Fermi-Hubbard model \cite{Pruschke:1993aa}. To assess these errors it is useful to check the accuracy to which spectral sum rules (valid also in a nonequilibrium setting \cite{Freericks:2013oq}) are fulfilled.
The moments $\mu^R_n(T)$ of the spectral function $A^R(T, \omega)$, $\mu^R_n(T) = \int_{-\infty}^\infty d\omega \, \omega^n A^R(T, \omega)$, are given by the higher order derivatives of the retarded Green's function $G^R(T, t)$ at $t=0^+$, $\mu^R_n(T) = - \textrm{Im} [ i^n \partial_t^n G^R(T, t) ]_{t=0^+}$, where $T$ and $t$ are the absolute and relative time respectively, see Ref.\ \onlinecite{Freericks:2013oq}. The moments can also be determined using the equation of motion, in terms of operator expectation values, $\mu^R_0 = 1$, $\mu^R_1 = \langle \epsilon \rangle -\mu + 2\langle \hat{n} \rangle U$, and $\mu^R_2 = \langle \epsilon^2 \rangle + \mu^2 + 3U^2\langle \hat{n}^2 \rangle - \langle \hat{n} \rangle(4\mu U + U^2)$, where $\langle \epsilon^n \rangle$ denotes the $n$th moment of the non-interacting density of states, see Ref.\ \onlinecite{Anders:2011uq}. For an approximate solution of the BDMFT equations, these approaches do not yield the same result.
In equilibrium, BDMFT+NCA gives a 1.6\% relative error of the first spectral moment $\mu^R_1$ in the Mott insulating phase ($U=96$, $T=6$), and a 10\% error in the vicinity of the superfluid phase boundary ($U=40$, $T=6$). For the second moment $\mu^R_2$, the relative errors are 11\% and 36\%, respectively.
\subsection{Quenches from the Mott insulator} \label{sec:Mott}
As a first application, we study quenches within the symmetric Mott and normal phases ($|\phi|=0$), i.e.\ suppressing symmetry breaking (superfluid states). Since the Gutzwiller mean-field description only contains the condensate energy $E_c$ and the interaction energy $E_i$, the symmetric state in this approximation is simply the atomic limit with $E_c \propto |\phi|^2=0$. So, for these quenches, no energy conversion occurs in the Gutzwiller treatment, resulting in an unphysical constant time-evolution.
BDMFT, however, retains temporal fluctuations and enables the conversion of interaction energy $E_i$ into kinetic energy $E_k$, and vice versa, which leads to a non-trivial quench dynamics.
To search for thermalization we study the relative change $\kappa$ in the double occupancy $\langle \hat{n}^2 \rangle$,
\begin{equation}
\kappa = \frac{\langle \hat{n}^2 \rangle(t) - \langle \hat{n}^2 \rangle_{U_i, T_i}}
{\langle \hat{n}^2 \rangle_{U_f, T_{\textrm{eff}}} - \langle \hat{n}^2 \rangle_{U_i, T_i}}
\label{eq:kappa} \, ,
\end{equation}
defined so that $\kappa(t=0) = 0$ and $\kappa = 1$ for a thermalized state.
Using this quantity, we identify an intermediate region of rapid thermalization in the $(U, T)$ plane [Fig.\ \ref{fig:UTPhaseDiagram}a] in the following way: For a given initial state $(U_i, T_i)$ we locate the maximum of $|1-\kappa(t=t_{\textrm{max}})|^{-1}$ as a function of $U_f$ at the longest accessible time $t_{\textrm{max}}=2.66$, as shown explicitly for $U_i=30$ and $T_i=6$ in Fig.\ \ref{fig:UTPhaseDiagram}b. The values of $U_f$ and the corresponding effective temperatures $T_\text{eff}$ are shown in Fig.\ \ref{fig:UTPhaseDiagram}a, and the result turns out to be insensitive to the initial interaction ($U_i = 30$, $45$). In both the weak and strong-coupling $U_f$ regimes the system is trapped in a long-lived ``prethermalized state'', reminiscent of the relaxation dynamics in the paramagnetic Fermi-Hubbard model \cite{Eckstein:2009fu}.
The observed absence of thermalization in these regimes can be understood in terms of proximity to an integrable point, since the Bose-Hubbard model is integrable both for $U=0$ and $U=\infty$ \cite{Sorg:2014vn}.
Interestingly the relaxation behavior in the two regimes differ. In the strong-coupling regime the exponential decay of $|1 - \kappa|$ slows down as $U_f$ increases (green, red, cyan lines in Fig.\ \ref{fig:UTPhaseDiagram}d). In the low $U_f$ regime $\kappa$ very rapidly reaches a plateau value (blue line in Fig.\ \ref{fig:UTPhaseDiagram}d), which increases roughly exponentially as $U_f$ is decreased.
The crossover between these two disparate behaviors is hard to pinpoint, and the indicated regions in Fig.\ \ref{fig:UTPhaseDiagram}a are only qualitative as $|1-\kappa(t=t_{\textrm{max}})|^{-1}$ is $t_{\textrm{max}}$ dependent (the region becomes narrower and shifts to slightly higher $U_f$ with increasing $t_{\textrm{max}}$).
\subsection{Quenches from the superfluid} \label{sec:SuperFluid}
\subsubsection{Relaxation regimes} \label{sec:SuperFluidRegimes}
\begin{figure}
\includegraphics[scale=1]{figure5.pdf}
\caption{\label{fig:SmallUf} (color online) Interaction quenches starting in the superfluid state ($U_i=6$, $T_i=5.1$), with
a) the positions of the final states ($U_f$, $T_{\textrm{eff}}$) (crosses) and the $(U, T)$ equilibrium superfluid phase boundary (gray line), and b) the corresponding evolution of the magnitude of the order parameter $|\phi|(t)$.
For $U_f=6.6$ the (equilibrium) thermal reference state at $(U_f, T_{\textrm{eff}})$ is also superfluid and the non-equilibrium dynamics displays a rapid transient growth of the condensate (cyan). Close to the phase boundary in the normal phase the system is trapped in a superfluid for long times and the quench induced excitation is transferred to an amplitude mode at longer times (magenta). For intermediate $U_f$ a constant trapped superfluid persists (blue) and after passing the dynamical transition $U_{c}^{\textrm{dyn}}$ the system undergoes exponential relaxation to the normal phase (red).
For large final interactions $U_f \gg W,zJ$ the condensate displays ``collapse-and-revival" oscillations with frequency $\omega \approx U_f$ (green). The evolution from $U_f=6.6$ to $U_f=11.4$ is also shown (gray lines).
For the growing condensate at $U_f=6.6$ panel c) shows the accuracy in total energy $E_t$ and density $\langle \hat{n} \rangle$ and panel d) shows the energy conversion during the time of rapid condensate growth.
} \end{figure}
The non-equilibrium dynamics after a quench from the superfluid ($|\phi| > 0$) with weak interaction $U_i = 6$ to larger interactions $U_f > U_i$ generates a variety of dynamical behaviors depending on $U_f$.
Apart from $U_f$ the system has two other characteristic energies (or inverse timescales), namely the bandwidth $W=12J$ and the condensate coupling $zJ = 6J$ (where $J=1$ is our unit of energy). In general the time evolution can be separated into five regimes, see Fig.\ \ref{fig:SmallUf}a and \ref{fig:SmallUf}b.
i) For quenches deep into the Mott regime, i.e.\ for large $U_f \gg W, zJ$, the condensate oscillates with the frequency $\omega$ of the final interaction strength, $\omega \approx U_f$, while relaxing exponentially (green line).
The relaxation rate strongly depends on the initial temperature $T_i$. For high $T_i$ (as in Fig.\ \ref{fig:SmallUf}b) the system displays relaxation to the Mott phase, while for low $T_i$ the system is trapped in a non-equilibrium superfluid state for long times, see Sec.\ \ref{sec:LongTimeDynamics}.
ii) In the intermediate coupling regime $U_f \gtrsim W > zJ$ the interaction driven oscillations compete with the kinetic time scale and only a few oscillations can be observed, and after the condensate time scale $2\pi/(zJ)$ the system displays exponential relaxation (red line).
iii) For $W \gtrsim U_f > zJ$ the thermal reference state is closer to the phase boundary in the normal phase. In this regime, after an initial transient undershoot in $|\phi|$, the system becomes trapped in a non-equilibrium superfluid state with a constant non-zero condensate (blue line).
iv) In the same range of $U_f$ an amplitude mode is excited at longer times (magenta line) with a roughly constant frequency but growing amplitude as the phase boundary is approached from the normal phase side.
v) For small $\Delta U = U_f - U_i$, ($W > U_f \approx zJ$) the initial transient is weak as the quench-energy scales with $\Delta U$. First the condensate undergoes a weak oscillatory transient followed by a sudden rapid growth (cyan line). This growth occurs when the final state is in the equilibrium superfluid region.
The rapidly growing condensate is a numerical challenge because of the occupation of high-energy (i.e.\ high occupation-number) states.
On the one hand, the cutoff in the bosonic Fock space must be chosen large enough to accommodate this, and on the other hand, the fast oscillations of the high-energy modes require a small time-step.
In Fig.\ \ref{fig:SmallUf}b We plot the results up to the point to which they can be fully converged both in the size of the time-step and the size of the Hilbert space. For $N_\text{max}=11$ and $\Delta t = 0.005$ the drift in total energy $\Delta E_t = |E_t(t) - E_t(0)|$ and density $\Delta \langle \hat{n} \rangle = |\langle \hat{n}(t) \rangle - \langle \hat{n}(0) \rangle|$ is of the order $\lesssim 10^{-6}$, see Fig.\ \ref{fig:SmallUf}c.
Hence, we conclude that the growth is a robust feature of our BDMFT+NCA calculations.
During the growth there is a rapid conversion between the different energy components of the system, while the total energy $E_t$ is conserved, see Fig.\ \ref{fig:SmallUf}d. The interaction energy $E_i$ and the normal component of the kinetic energy $E_{k}^{(n)} \propto \langle b_i^\dagger b_{i+1} \rangle$ rapidly increase, while the condensate energy $E_c$ and the anomalous component of the kinetic energy $E_{k}^{(a)} \propto \langle b_i b_{i+1} \rangle$ decrease by the same total amount.
The self-amplified transient growth of the condensate fraction resembles the quantum turbulence driven dual cascade with non-equilibrium Bose-Einstein condensation observed in scalar field theories \cite{Berges:2012uq}.
Also in other contexts, dynamical instabilities with diverging solutions have been observed in lattice boson systems. In the weak coupling limit, a Gross-Pitaevskii treatment yields dynamically unstable solutions \cite{Machholm:2003ly}, also observed experimentally \cite{De-Sarlo:2005kl}. In the Bose-Hubbard model the exponential condensate growth of symmetric initial states has previously been studied in mean-field \cite{Snoek:2011hc}.
However, for the weak interaction quenches within the superfluid we cannot rule out that the sudden condensate growth is an artifact of the NCA treatment. Higher-order implementations of the self-consistent strong coupling expansion may in the future help to clarify this issue.
\subsubsection{Short time dynamics
after quenches deep into the Mott regime} \label{sec:ShortTimeDynamics}
\begin{figure}[b]
\includegraphics[scale=1]{figure6.pdf}
\caption{\label{fig:TiSweep} (color online) Superfluid quench short time dynamics for $U_i=6$ and $U_f=48$ deep within the Mott phase ($U_f \gg W > zJ$). Upper panel: The first revival maximum coincides with the final interaction period $n \cdot 2\pi/U_f$ (black dotted lines). Exponential fits for the relaxation, using the first revival maximum at $t = 2\pi/U_f$ (colored dotted lines) are also shown. Lower panel: Time dependence of $(\partial_t \theta) |\phi|^2$.
} \end{figure}
\begin{figure*}
\includegraphics[scale=1]{figure7.pdf}
\caption{\label{fig:SFquench} (color online) a) Interaction quenches from superfluid initial states with $U_i=6$, $\langle \hat{n} \rangle = 1$, and initial temperatures $T_i = 4.5$ (magenta) and $5.1$ (yellow), close to $T_c \approx 5.49$.
b) The window averages $\bar{|\phi|}(t=t_{\textrm{max}})$ are suppressed in the crossover region (red shaded region). c) Typical time evolutions of $|\phi|$ (solid lines) and $\bar{|\phi|}(t=t_{\textrm{max}})$ (dotted lines), for $T_i = 4.5$, in the three regimes and d) mean-field results for the same parameters are also shown.
e) The phase frequency $\omega_{\theta}$ exhibits a kink (arrows) at the dynamical transition $U_{c}^{\textrm{dyn}} \approx 13.5$ and $18.0$ where f) the double occupancy relaxation $\tau^{-1}_{\kappa}$ is peaking and the damping of the amplitude-mode $\tau^{-1}_{|\phi|_{AM}}$ is maximal, while g) the condensate amplitude relaxation $\tau^{-1}_{|\phi|_c}$ peaks after $U_{c}^{\textrm{dyn}}$ (arrows).}
\end{figure*}
In the limit of large final interaction $U_f \gg W$, $zJ$, i.e.\ in regime (i), the superfluid quenches from $U_i = 6$ display oscillations with a frequency $\omega$ scaling with $U_f$, $\omega \approx U_f$, and the short time behavior is dominated by the interaction, as it defines the shortest time scale of the system.
The short time relaxation is expected to be driven by local decoherence and the long time relaxation ($t > 2\pi/W \approx 0.52$) to be dominated by hopping.
In order to study the short time dynamics we perform a series of quenches to $U_f = 48$ for several initial temperatures $T_i = 3.00, \dots , 5.25$, see Fig.~\ref{fig:TiSweep}.
While $\omega$ scales with $U_f$ there are important contributions from other frequency components. Pure $2\pi/U_f$-oscillations can only be observed in the first few revivals, while they are at later times washed out by the off-diagonal mixing of local occupation number states in the initial state.
The first revival maximum occurs at the period of the final interaction $2\pi / U_f$, the second revival has a pronounced two peak structure with the first peak occurring at $2 \cdot 2\pi/U_f$, and in the third revival the $3 \cdot 2\pi/U_f$ peak only appears as a shoulder of the main peak.
From Fig.\ \ref{fig:TiSweep} it is also evident that the short time decoherence strongly depends on the initial temperature $T_i$, with higher temperature resulting in faster damping.
An interesting question is whether the long time relaxation rate can already be inferred from the short time decoherence, in the spirit of the strong coupling analysis of Ref.\ \onlinecite{Fischer:2008ys}. To investigate this we fit the simple exponential model $|\phi|(t) \approx |\phi|(0) e^{-t / \tau}$ to the real time data, where the relaxation rate $\tau$ is approximated using $|\phi|$ at the first revival maximum $t_r = 2\pi/U_f$ as $\tau = t_r / \log(|\phi|(t_r) /|\phi|(0))$.
Figure \ref{fig:TiSweep} shows that the relaxation rate $\tau$ is overestimated for low temperature initial states and underestimated for high temperature initial states.
Hence in this regime the condensate relaxation can not be inferred from the first revival maximum.
Infact the long time exponential relaxation rate is only established after the characteristic condensate time scale $2\pi/(zJ)$, see e.g.\ the green and red lines in Fig.~\ref{fig:SmallUf}b.
We also note that the BDMFT calculation does not involve any approximation concerning the timescales of the dynamics. This sets it apart from, for example, the low frequency approximation applied in the Schwinger-Keldysh generalization of the strong coupling approach \cite{Kennett:2011fk}, where in the particle-hole symmetric limit $\langle \hat{n} \rangle = 1$ the condensate phase $\theta$ and amplitude $|\phi|$ (where $\phi = |\phi|e^{i\theta}$) are constrained by $(\partial_t \theta) |\phi|^2 = C$ for some constant $C$. As shown in the lower panel of Fig.~\ref{fig:TiSweep}, the BDMFT dynamics has a non-trivial time dependence in this quantity.
\subsubsection{Long time dynamics}
\label{sec:LongTimeDynamics}
The long time dynamics for quenches from the superfluid has been investigated in a number of zero temperature Gutzwiller mean-field studies \cite{Huber:2007ys, Wolf:2010bh, Sciolla:2010uq, Sciolla:2011kx, Snoek:2011hc, Krutitsky:2011qf}. The most prominent nonequilibrium effect is a dynamical transition at $U_f=U_{c}^{\textrm{dyn}}[U_i]$ \cite{Sciolla:2010uq}.
It is important to note, however, that for low $U_i$ only the mean-field calculation using a constrained basis including the lowest three bosonic occupation number states ($N_{\textrm{max}}= 3$) produces a sharp transition. If the physically important states with higher occupations are also considered, the transition turns into a crossover
\footnote{Note that the $N_{\textrm{max}} > 3$ results of Refs.\ \onlinecite{Sciolla:2010uq} and \onlinecite{Sciolla:2011kx} are in the high-$U_i$ regime.}.
In a broader perspective the occurrence of a dynamical transition is not specific to the Bose-Hubbard model, but has also been observed in the Fermi-Hubbard model and other systems on the mean-field level \cite{Schiro:2010fk, Gambassi:2011dz, Sciolla:2011kx}.
The quantum fluctuations missing in mean-field treatments are expected to heavily modify the dynamical transition, as previously shown for other systems, using dynamical mean field theory \cite{Eckstein:2009fu}, the Gutzwiller approximation including Gaussian fluctuations \cite{Schiro:2011oq, Sandri:2012fv}, and $1/N$ expansions \cite{Sciolla:2013bs}.
Here we show how the dynamical transition in the Bose-Hubbard model is affected when we go beyond the simple mean-field treatment, starting from a thermal initial state, and include quantum fluctuations using BDMFT.
As the hopping induced relaxation is most prominent for small $U_i$ and temperatures $T_i$ close to the superfluid phase boundary, we fix $U_i=6$, far away from the zero temperature transition, $U_c(T=0) \approx 26.4$, see Fig.~\ref{fig:PhaseDiag}. To see the enhanced relaxation in the vicinity of the phase boundary, located at $T_c(U=6) \approx 5.49$, we consider the two initial temperatures $T_i = 4.5$ and $5.1$ with relatively weak superfluidity $|\phi|^2 \lesssim 0.5$, see Fig.\ \ref{fig:SFquench}a.
To analyze the dynamics of the condensate amplitude $|\phi|$
we first look at windowed time averages $\bar{|\phi|}(t)$, using a Gaussian window with width $t_w = 1/3$ to filter out oscillations.
In Fig.\ \ref{fig:SFquench}b we plot the window average $\bar{|\phi|}(t=t_{\textrm{max}})$ at the longest time as a function of $U_f$, thereby restricting ourselves to the regimes (i)-(iii) above, i.e., when the final equilibrium state is not in the superfluid phase and the order parameter does not show self-amplified growth.
From Fig.\ \ref{fig:SFquench}b it is evident that $\bar{|\phi|}(t=t_{\textrm{max}})$ exhibits a crossover for $T_i=4.5$, from high values at low $U_f$ close to $U_i=6$, through a minimum at intermediate $U_f$, and increasing again for $U_f \gtrsim 30$, in qualitative agreement with the mean-field dynamical transition \cite{Sciolla:2010uq}. Also the general temperature dependence, namely that a higher temperature leads to lower condensate averages, agrees with mean-field.
However the thermal effects in BDMFT are much stronger: both the minimum and the large $U_f$ plateau are drastically reduced, going from $T_i=4.5$ to $T_i=5.1$. As we will show, this reduction is due to a rapid condensate relaxation rate $\tau^{-1}_{|\phi|_c}$ emerging close to the phase boundary (Fig. \ref{fig:SFquench}g).
The BDMFT real-time evolution in the three regimes is shown in Fig.\ \ref{fig:SFquench}c. For small $U_f$ in regime (iii), $|\phi|$ stabilizes at a finite value after an initial transient, even though the thermal reference state is in the normal phase. The intermediate regime (ii) shows fast thermalization with a rapidly decaying condensate and damped collapse-and-revival oscillations, in qualitative agreement with the experimental results of Greiner {\it et al.~}\cite{Greiner:2002fk}. Interestingly, for larger $U_f$ in regime (iii) the system is again trapped in a nonthermal superfluid phase, now exhibiting coherent amplitude oscillations with finite life-time.
Note that none of these relaxation and thermalization effects are captured by the Gutzwiller mean-field approximation, which predicts an oscillatory behavior (Fig.~\ref{fig:SFquench}d).
While the minimum of $\bar{|\phi|}(t=t_{\textrm{max}})$ indicates a crossover, the dynamical transition $U_{c}^{\textrm{dyn}}$ can be accurately located by studying the condensate phase $\theta(t)$, where $\phi(t) = |\phi|(t) \, e^{i\theta(t)}$.
Its linear component $\partial_t \theta(t) \approx \omega_{\theta}$ exhibits a kink at $U_f = U_{c}^{\textrm{dyn}}$, see Fig.\ \ref{fig:SFquench}e, in direct analogy with mean-field predictions where $\theta$ (mapped to a conjugate momentum $p$) also has a slope discontinuity at $U_{c}^{\textrm{dyn}}$ \cite{Sciolla:2010uq}.
Similar to the fast thermalization region in the symmetric phase, the double occupancy thermalizes rapidly for $U_f \approx (U_{c}^{\textrm{dyn}})^+$ as can be seen in Fig.\ \ref{fig:SFquench}f from the drastic increase in the relaxation $\tau_{\kappa}^{-1}$ of
\begin{equation*}
|1-\kappa| \propto e^{-t / \tau_{\kappa}} \, ,
\end{equation*}
as $U_f \rightarrow (U_{c}^{\textrm{dyn}})^+$.
To get a qualitative understanding of the condensate amplitude $|\phi|$ relaxation dynamics, we fit the late time-evolution $(t>1.3)$ to a damped two component model
\begin{equation*}
|\phi_M|(t) =
A_{|\phi|_c} e^{-t/\tau_{|\phi|_c}} + A_{|\phi|_{AM}} \cos^2(\omega t+\varphi) e^{-t/\tau_{|\phi|_{AM}}},
\end{equation*}
with a non-oscillatory component (${|\phi|_c}$) and a coherent amplitude-mode (${|\phi|_{AM}}$), and relaxation $\tau^{-1}_{|\phi|_c}$ and damping $\tau^{-1}_{|\phi|_{AM}}$ respectively, see Fig.\ \ref{fig:SFquench}f and \ref{fig:SFquench}g.
The amplitude-mode frequency $\omega$ has the same general behavior as $\omega_{\theta}$ (not shown), and $\omega, \omega_{\theta} \rightarrow U_f$ in the large $U_f$ limit.
Analogous to the rapid relaxation of the double occupancy, the amplitude mode is strongly damped for $U_f \approx (U_{c}^{\textrm{dyn}})^+$, but it retains a finite lifetime $\tau^{-1}_{|\phi|_{AM}} > 0$ for large $U_f$, see Fig.\ \ref{fig:SFquench}f. The relaxation of the non-oscillatory component shows two distinct behaviors: For $U_f < U_{c}^{\textrm{dyn}}$ the system is trapped in a superfluid state and the condensate relaxation is almost zero, $\tau_{|\phi|_c}^{-1} \approx 0$, while for $U_f \gtrsim U_{c}^{\textrm{dyn}}$ it becomes finite, reaching a maximum at intermediate $U_f$, see Fig.\ \ref{fig:SFquench}g.
For large $U_f$ and $T_i=5.1$, $\tau_{|\phi|_c}^{-1}$ stays finite and the system eventually thermalizes to the Mott state, while for $T_i=4.5$, $\tau_{|\phi|_c}^{-1}$ becomes small as $U_f \rightarrow \infty$, which means that the system is trapped for a very long time in a superfluid state.
The stability of the superfluid can be understood in terms of a simple two fluid model of doublons and hard-core bosons \cite{Sorg:2014vn}. In this picture the quench generates long-lived doublons and depletes the hard-core boson gas away from unity filling, where it can remain a superfluid for any local interaction \cite{Schneider:2014ve}.
This case is particularly interesting as it opens up the possibility to study the Higgs amplitude mode in a metastable superfluid.
\subsubsection{Nonequilibrium ``phase diagram"}
\label{sec:SuperFluidPhaseDiagram}
\begin{figure}
\includegraphics[scale=1]{figure8.pdf}\caption{\label{fig:PhaseDiagram} (color online)
Qualitative non-equilibrium $(U_f, T_{\textrm{eff}})$ phase diagram for quenches from equilibrium states (crosses) in the superfluid $(U, T)$ phase region (gray), with $U_i = 6$ and initial temperatures $T_i = 3.50$, $4.00$, $4.50$, $5.10$, and $5.25$, to final states with $(U_f, T_{\textrm{eff}})$ (circles) within the equilibrium normal phase region.
The boundary between the trapped superfluid (SF) (blue) and the crossover region (red) is given by $U_{c}^{\textrm{dyn}}$ (squares), and the boundary between the crossover (red) and trapped superfluid with amplitude mode (SF+AM) (green) is taken as the point where $\tau^{-1}_{|\phi|_c}$ drops below $50\%$ of its crossover peak-value on the high-$U_f$ side (triangles), see Fig.~\ref{fig:SFquench}g.
} \end{figure}
We summarize the results for the long-time dynamics in the non-equilibrium ``phase diagram" shown in Fig.~\ref{fig:PhaseDiagram}.
By repeating the analysis for $U_i=6$ and the series $T_i = 3.50$, $4.00$, $4.50$, $5.10$, and $5.25$ of initial temperatures, we locate the boundaries of the three dynamical regimes in the equilibrium normal phase region, regime (i), the high-$U$ region characterized by a trapped superfluid and amplitude mode (green), regime (ii), the crossover region with rapid thermalization (red), and regime (iii), the trapped superfluid in the vicinity of the equilibrium superfluid phase (blue).
While this non-equilibrium ``phase diagram'' depends on the initial states and the quench protocol, it gives an overview of the different relaxation and trapping phenomena and their location in parameter space.
An experimental verification of these different dynamical regimes of the Bose-Hubbard model would be very interesting and presumably possible.
In the one dimensional Bose-Hubbard model a similar behavior has been theoretically observed using DMRG \cite{Kollath:2007ys}, and for longer times using time-dependent variational Monte Carlo \cite{Carleo:2012zr}. Quenches from the zero temperature superfluid display a region of thermalization at intermediate final interactions and a trapping in nonthermal states for long times at strong final coupling \cite{Kollath:2007ys}. At unity filling this behavior can be understood in terms of a reduced effective scattering of holon and doublon excitations at strong interactions \cite{Altman:2002ve}. Our results from BDMFT indicate that this phenomenon is also relevant in three dimensions (green region in Fig.\ \ref{fig:PhaseDiagram}). However, we also identified a transient trapping at low interactions (blue region in Fig.\ \ref{fig:PhaseDiagram}) which has not been reported for 1D. It is an open question whether this feature is specific to high-dimensional models.
We also note that while BDMFT allows us to compare nonequilibrium and equilibrium states within the same formalism, the DMRG studies involve comparisons between time-dependent correlators and finite-temperature QMC results \cite{Kollath:2007ys}.
\section{Conclusions} \label{sec:Conclusions}
We developed the nonequilibrium BDMFT formalism and its implementation in combination with a NCA type bosonic impurity solver. We have demonstrated its ability to capture nontrivial dynamical effects in quenched Bose-Hubbard systems, including dynamical transitions, fast-thermalization crossovers, and trapped superfluid phases with long-lived but damped amplitude oscillations. These results were collected into two nonequilibrium ``phase diagrams'' (Figs.\ \ref{fig:UTPhaseDiagram} and \ref{fig:PhaseDiagram}), which illustrate the transitions and crossovers that occur as one varies the quench parameters. Particularly noteworthy results are the prediction of a very long-lived transient superfluid state with an amplitude mode after quenches from the superfluid phase into the Mott regime, our finding of a trapped superfluid state after quenches into the vicinity of the superfluid phase boundary, and the nonequilibrium Bose condensation (growing condensate) after small quenches within the superfluid phase.
The ability of BDMFT to describe hopping induced relaxation phenomena at finite temperature goes beyond all current competing theoretical approaches. The Gutzwiller mean-field formalism lacks all hopping induced phenomena \cite{Sciolla:2010uq, Sciolla:2011kx}, and the strong coupling based real-time approach \cite{Kennett:2011fk} is limited to zero temperature and slow dynamics. The hopping perturbation expansion \cite{Trefzger:2011uq} looks promising but has so far only been applied at zero temperature. A comparative study with the finite temperature extension of this approach would be very interesting.
Extensions of the nonequilibrium BDMFT formalism to multi-flavor Hamiltonians \cite{Soyler:2009ys,Capogrosso-Sansone:2010zr} and inhomogeneous systems (e.g.\ with a trapping potential) \cite{Eckstein:2013fv} should enable direct comparisons with cold atom experiments.
Multi-orbital effects such as virtual excitations to higher orbitals can trivially be included in BDMFT in terms of effective three body interactions \cite{Johnson:2009kx}.
While calculations based on unitary-time evolution \cite{Tiesinga:2011ve} suffice to understand experiments \cite{Will:2010uq} in the $J \rightarrow 0$ limit, BDMFT can extend the theoretical treatment to finite $J$.
An inhomogeneous extension of BDMFT will require a more advanced parallelization scheme than that applied by Dirks et al.\ \cite{Dirks:2014ij} for the fermionic case, but it would enable studies of very important phenomena, such as mass transport and the effects of the trapping potential in general cold-atom systems out-of-equilibrium.
The big challenge in these systems is the inherent disparity of the hopping and mass transport time-scales.
Simpler approximations such as the hopping expansion has successfully been applied in this context \cite{Dutta:2014uq},
but without incorporating thermal and retarded correlation effects.
In a broader perspective it should be productive to apply nonequilibrium BDMFT or variants of this formalism to nonequilibrium Bose-condensation in, e.g., polaritonic systems and field theories \cite{Carusotto:2013kx, Berges:2012uq, Akerlund:2013fk}.
\begin{acknowledgments}
The authors would like to acknowledge fruitful discussions with
H. Aoki,
T. Ayral,
J. Berges,
N. Buchheim,
D. Golez,
F. Heidrich-Meisner,
A. Herrmann,
S. Hild,
D. H\"{u}gel,
M. P. Kennett,
Y. Murakami,
L. Pollet,
U. Schneider,
N. Tsuji,
and
L. Vidmar.
The calculations have been performed on the UniFr cluster. HS and PW are supported by FP7/ERC starting grant No.\ 278023.
\end{acknowledgments}
\section{Nambu generalization of the non-crossing approximation}
\label{app:NCA}
The solution of impurity actions without symmetry breaking by means of self-consistent strong-coupling perturbation theory, i.e.\ the non-crossing approximation (NCA) and its higher-order generalizations, has been discussed in detail in Ref.\ \onlinecite{Eckstein:2010fk}.
To apply this method to the BDMFT action in Eq.\ (\ref{eq:Seff}) we have to extend the NCA formalism to Nambu spinors and symmetry broken states.
The diagrammatics of Ref.\ \onlinecite{Eckstein:2010fk} needs to be modified on the operator and hybridization function level.
While the pseudo-particle propagators $\hat{G}_{\Gamma \Gamma'}(t, t')$ still only carry local many-body state indices $\Gamma$ and $\Gamma'$ (corresponding to the occupation number states $| \Gamma \rangle$ and $| \Gamma' \rangle$), the hybridization function $\boldsymbol{\Delta}_{\gamma \nu}(t, t')$ now carries two Nambu indices $\gamma$ and $\nu$.
We will represent the propagators with directed solid and dashed lines according to
\begin{equation*}
\includegraphics{appEqProp.pdf}
\end{equation*}
where $t$ and $t'$ are times on the contour $\mathcal{C}$.
Due to the Nambu indices of $\boldsymbol{\Delta}$, the vertices of the theory must also be equipped with a Nambu index $\gamma$, in combination with a contour time $t$ and in and out going many-body state indices $\Gamma'$ and $\Gamma$, respectively. The matrix elements can be graphically represented as
\begin{equation*}
\includegraphics{appEqVert.pdf}
\end{equation*}
where the direction of the hybridization line determines the operator of the vertex. A hybridization line entering a vertex creates a ``Nambu-particle'' by insertion of $\mbf{b}^\dagger_\gamma$ giving the matrix element $\langle \Gamma | \mbf{b}^\dagger_\gamma | \Gamma' \rangle$ , and an interaction line leaving the vertex annihilates a ``Nambu-particle'' through $\mbf{b}_\gamma$ giving $\langle \Gamma | \mbf{b}_\gamma | \Gamma' \rangle$. In the following we will use the operator symbols $\mbf{b}^\dagger_\gamma$ and $\mbf{b}_\gamma$ to represent these matrix elements as they act in the same Fock-space as the pseudo-particle propagator $\hat{G}$ and the pseudo-particle self-energy $\hat{\Sigma}$.
\subsection{Pseudo particle self-energy}
\label{app:NCASigma}
Following the diagram rules of Ref.\ \onlinecite{Eckstein:2010fk} the pseudo-particle self-energy $\hat{\Sigma}$ at first order in $\boldsymbol{\Delta}$, corresponding to the non-crossing approximation (NCA), takes the form of shell-diagrams
\begin{equation*}
\includegraphics{appEqSigma.pdf}
\end{equation*}
where $\xi = \pm 1$ for bosons and fermions respectively.
Using the Nambu generalization of propagators and vertices gives the contour expression for the first diagram with a forward propagating hybridization line according to
\begin{multline} \\[-8mm]
\hat{\Sigma}^{(1)}(t, t') = (i)
\includegraphics[valign=c]
{appEqSigmaFwd.pdf}
\\ =
\frac{i}{2} \sum_{\gamma \nu } \boldsymbol{\Delta}_{\gamma \nu}(t, t')
\big[ \mbf{b}^\dagger_\gamma(t) \, \hat{G}(t, t') \, \mbf{b}_\nu(t') \big]
\, , \\[-8mm]
\label{eq:NCASigma1}
\end{multline}
with implicit matrix multiplications in the many-body state indices $\Lambda$ and $\Lambda'$. The second diagram is constructed analogously
\begin{multline} \\[-8mm]
\hat{\Sigma}^{(2)}(t, t') =
(i\xi)
\includegraphics[valign=c]
{appEqSigmaBwd.pdf}
\\ =
\xi \frac{i}{2} \sum_{\gamma \nu} \boldsymbol{\Delta}_{\nu \gamma}(t', t)
\big[ \mbf{b}_\gamma (t) \, \hat{G}(t, t') \, \mbf{b}^\dagger_\nu(t') \big]
\, . \\[-8mm]
\label{eq:NCASigma2}
\end{multline}
Suppressing many-body state indices in these diagrams yields Fig.\ \ref{fig:NCA}a in Sec.\ \ref{sec:NCA}. Collecting all terms, we obtain
\begin{align*}
\hat{\Sigma}(t, t') =
& \frac{i}{2} \sum_{\gamma \nu}\boldsymbol{\Delta}_{\gamma \nu}(t, t')
\big[ \mbf{b}^{\dagger}_\gamma(t) \, \hat{G}(t, t') \,\mbf{b}^{\phantom{\dagger}}_\nu(t') \big]
\\ +
\xi & \frac{i}{2}\sum_{\gamma \nu}\boldsymbol{\Delta}_{\nu \gamma}(t', t)
\big[ \mbf{b}^{\phantom{\dagger}}_\gamma(t) \, \hat{G}(t,t') \,\mbf{b}^{\dagger}_\nu(t') \big]
\, ,
\end{align*}
which corresponds to Eq.\ (\ref{eq:NCASigma}) in Sec.\ \ref{sec:NCA}.
To perform actual calculations we work with a subset of Keldysh components \cite{Aoki:2014kx}, namely,
the imaginary time Matsubara component $\hat{\Sigma}^M(\tau)$,
the real-time greater component $\hat{\Sigma}^>(t, t')$,
the real-time lesser component $\hat{\Sigma}^<(t, t')$,
and the right-mixing component $\hat{\Sigma}^\urcorner(t, \tau')$.
These components can be derived from the general contour expression for $\hat{\Sigma}$ using the Langreth product rules \cite{Eckstein:2009qf, Stefanucci:2013oq}.
This is because the pseudo-particle self energy $\hat{\Sigma}$ is given by contour time products of the hybridization function $\boldsymbol{\Delta}$ and the pseudo-particle propagator $\hat{G}$, $\hat{\Sigma} \propto \boldsymbol{\Delta} \hat{G}$. Note that the two contributions $\hat{\Sigma}^{(1)}$ and $\hat{\Sigma}^{(2)}$ [Eqs.\ (\ref{eq:NCASigma1}) and (\ref{eq:NCASigma1})] must be treated differently as the order of the time arguments in $\boldsymbol{\Delta}$ differs.
\subsection{Single particle Green's function}
\label{app:NCAGsp}
The NCA approximation for the single particle Green's function is given by the pseudo-particle bubble equipped with two vertices \cite{Eckstein:2010fk}.
The Nambu generalization amounts to adding Nambu indices to the vertices and gives the non-connected Green's function as
\begin{multline}
\tilde{\mbf{G}}_{\gamma \nu}(t, t') = (i) \times
\textrm{Tr} \Bigg[
\includegraphics[valign=c]
{appEqGf.pdf}
\Bigg]
\\ =
i \textrm{Tr} \big[ \hat{G}(t', t) \,\mbf{b}^{\phantom{\dagger}}_\gamma (t) \, \hat{G}(t, t') \,\mbf{b}^{\dagger}_\nu(t') \big]
\, ,
\end{multline}
where the hybridization line stubs denote the insertion of a Nambu creation or annihilation operator, and the trace corresponds to the summation over the $\Gamma'$ many-body state index.
To obtain the connected Green's function $\mbf{G}$ from $\tilde{\mbf{G}}$, the symmetry broken contribution must be removed, i.e.\ $\mbf{G}(t, t') = \tilde{\mbf{G}}(t, t') + i \Phi(t) \Phi^\dagger(t')$, which corresponds to Eq.\ (\ref{eq:NCAGsp}) and Fig.\ \ref{fig:NCA}b in Sec.\ \ref{sec:NCA}.
|
\section{The model}
Let us consider the following Hamiltonian,
\begin{eqnarray}
H&=&H_S+\sum_k \omega(k) a_k^\dagger a_k+\sum^N_{n=1}g_n\int_{-k_s}^{k_s}dk u_n (a_k^\dagger \phi_n(k)\cr
&+& a_k \phi^*_n(k)) C_n.
\label{h1EM}
\end{eqnarray}
with $k_s$ the maximum wave vector $k$ of the environment, $N$ the number of particles in the quantum system, $a^\dagger_k (a_k)$ the creation (annihilation) operators corresponding to the environment mode $k$ with frequency $\omega(k)$, $u_n$ the coupling strength of the system particle $n$ with such field, and $C_n$ a system coupling operator corresponding to the particle $n$. Also, $H_S$ is the system free Hamiltonian, and the functions $\phi_n(k)$ are here assumed to form an orthonormal set with the property (i) $\int_{-k_s}^{k_s}dk\phi^*_n(k)\phi_m(k)=\rho_0\delta_{nm}$, with $\rho_0$ a normalization factor.
Let us now consider the transformation $b_n=\int_{-k_s}^{k_s}dk \phi^*_n(k)a_k$, with the inverse
$a_k=\sum_m \phi_m(k)b_m$. To ensure that the conmutation (or anticommutation) properties of the original operators are conserved in the transformed ones, $b_n$,
we need an additional property (ii) $\sum_n \phi^*_n(k)\phi_n(k')=\rho_0\delta(k-k')$. Properties (i)-(ii) guarantee that the transformation is unitary.
Applying this transformation to (\ref{h1EM}) we obtain
\begin{eqnarray}
H=H_S+\sum^M_{n,m=1}f_{nm} b_n^\dagger b_m+g\sum^N_{n=1} u_n (b_n^\dagger+b_n)C_n,
\label{fintrans}
\end{eqnarray}
where we have assumed that $g_n=gu_n$, and $f_{nm}=\int_{-k_s}^{k_s}dk \omega_k\phi^*_n(k)\phi_m(k)$ is a function that decays with $|n-m|$ depending on the particular dispersion relation, and the orthonormal basis choosen. The original Hamiltonian has been mapped to that of a chain of transformed modes $b_n$ coupled one by one to each particle $n$. Note that although the proposed transformation
may require a large number $M$ of modes $b_n$ to correctly map the environment, the number of particles $N$ in the OQS can in principle be arbitrary.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.5\textwidth]{ChainChin2.pdf}}
\caption{(Color online) (a) Several atoms uniformly coupled to the same environment, according to the proposal by \cite{chin2013}. (b) Ladder structure here proposed, corresponding to the mapped Hamiltonian (\ref{fintrans}). We have denoted $g_n=gu_n$. \label{statNM}}
\end{figure}
The mapping is valid for both bosonic and fermionic reservoirs, and constitutes an interesting playground to analyze the dynamics of many body open quantum systems.
Furthermore, it can be used to describe the dynamics of many physically realistic problems. To illustrate this, in the following we analyze how light-matter interaction Hamiltonians can be tailored to reach the desired form (\ref{h1EM}), from which the proposed mapping can be performed.
\section{One dimensional electromagnetic fields}
Let us consider a Hamiltonian of the form
\begin{eqnarray}
H&=&H_S+\sum_k \omega(k)a_k^\dagger a_k+\sum^N_{n=1}\sum_k g_n(k)(a_k^\dagger u_k(r_n)\cr
&+& a_k u^*_k(r_n)) C_n,
\label{h1}
\end{eqnarray}
with
\begin{eqnarray}
g_{n}(k)=-i \sqrt{\frac{1}{2\hbar \omega(k) \epsilon_0\nu}}\omega_{n}\sum_\sigma\hat{e}_{{\bf k},\sigma}\cdot{\bf d}^n_{12},
\label{chapuno23}
\end{eqnarray}
where $\nu$ is the quantization volume, $\epsilon_0$ is the free space permitivity, and ${\bf d}^n_{12}=d^n_{12}{\bf u}_d$ and $\omega_n$ are respectively the dipole moment and resonant frequency of the $n$-th atom. In the following, we assume for simplicity that all atoms have equal resonant energy $\omega_0$ and dipolar moment $d_{12}$. The quantity $\hat{e}_{{\bf k},\sigma}$, is the polarization vector corresponding to the wave vector ${\bf k}$ with polarization $\sigma$. Here we have assumed that the energy absorption and emission process is independent on the polarization of the photons, i.e. $a_{k,\sigma}\equiv a_k$.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.5\textwidth]{ChainEM1D.pdf}}
\caption{(Color online) Ladder structure for $P=2$. The fact that the atoms are separated by a distance $d_0=2h_0$ gives rise to a situation where only one every two sites in the harmonic oscillator lattice is connected with an atom. \label{statNM2}}
\end{figure}
The functions $u_k(r_n)$ are the so-called mode functions, that in the continuum form an orthonormal set within the physical region under consideration, i.e. $\int dr u^*_k(r)u_k(r)=\delta_{kk'}$. These functions are solutions of the wave equation, and depend on the boundary conditions considered: while periodic boundary conditions correspond to travelling-wave modes, reflecting walls lead to standing waves \cite{wallsbook}. For the earlier case, $u_k(r_n)=(1/L)e^{ikr_n}$, and the wave vector $k$ takes the values $k=2\pi q/L$, with $L=h_0M$ defining the physical quantization region, and $q=1,\cdots,M$.
Naturally, to obtain this Hamiltonian from (\ref{h1EM}), we may just consider $\phi_n=h_0u_k(r_n)$, and assume that $g(k)\approx g$.
In order to fulfill the relation (ii) $(1/M)\sum_n\phi^*_n(k)\phi_m(k')=\delta(k-k')$, we shall chose atoms with positions $r_n=nd_0$, i.e. spaced by a constant distance given by $d_0=Ph_0$, where $P=1,2,\cdots$. In this situation, the proposed transformation becomes simply a discrete Fourier transform, and the mapped Hamiltonian has the form
\begin{eqnarray}
H=H_S+\sum^M_{n,m=1}f_{nm} b_n^\dagger b_m+\sum^N_{n=1} g_n (b_{nP}^\dagger+b_{nP})C_n,
\label{fintrans2}
\end{eqnarray}
Fig. (\ref{statNM2}) represents this structure for $P=2$. The larger the separation between atoms, $P$, the more spacing between the harmonic oscillators that are coupled to atoms within the chain.
Periodicity in the boundary conditions is not as restrictive as it seems: in general $N\ll M$, and atoms from $n=N+1$ to $n=M$ are just virtual particles introduced for the transformation. Hence, boundary effects do not affect drastically the systems dynamics, as long as the system is sufficiently small as compared to the environment.
Interesting at this point is to realize that when the initial state of the environment is thermal, the so-called correlation function is the only quantity necessary to describe the dynamics of an OQS coupled to it \cite{chin2011}. An exact calculation of the Heisenberg evolution equations of the system operators \cite{alonso2007}, leads to the conclusion that indeed the correlation function is the only quantity needed to describe the effects of the environment in the system dynamics (see Appendix A for details). Also, the correlation function appears in master equations derived with different methods \cite{breuerbook}. Hence, replacing $g_n(k)\approx g_n$ (with $g=g(k_\textmd{eff})$, and $k_\textmd{eff}$ is the resonant wave-vector, that can be defined as $\omega(k_\textmd{eff})=\omega_0$) in (\ref{h1}) is a reasonable approximation when in the calculus of the correlation function
\begin{eqnarray}
\alpha^{1D}_{nm}(t)=\frac{h_0}{2\pi}\int^{k_\textmd{m}}_{-k_\textmd{m}} dk g_n(k)g_m(k) e^{ikr_{nm}-i\omega(k)t},
\label{correl1D}
\end{eqnarray}
the term $\exp(i\omega(k)t)$ varies much faster in $k$ than $g(k)$, which can then be considered as a constant in comparison \cite{devega2005}. In other words, in many problems the structure of the environment is captured mainly by the dispersion relation, which in turn fully determines the density of states.
In the mapped Hamiltonian (\ref{fintrans}), in general $f_{nm}$ decays exponentially for low-order polynomial $\omega_k$, so provided that $\omega(k)$ is sufficiently smooth in $k$, $f_{nm}$ can be truncated at near or next-near neighbors.
This truncation is exact for periodic $\omega_k$ as we will see in Section V.
\section{Three dimensional EM fields}
When the many body open quantum system is connected to a vector field within a three dimensional space, the initial Hamiltonian may have a different form from (\ref{h1EM}). This is the case of atoms coupled to the electromagnetic field, where $H=H_S+H_B+H_\textmd{int}$, where $H_B=\sum_{\bf k}\omega({\bf k})a^\dagger_{\bf k}a_{\bf k}$. Considering the case of travelling waves, such that the magnitude of the mode functions is $u_{\bf k}({\bf r})=L^{-3/2}e^{i{\bf k}\cdot{\bf r}}$, the interaction Hamiltonian can be written as
\begin{eqnarray}
H^\textmd{3D}_\textmd{int}=\sum_{\bf k}\sum_{\bf n}g({\bf k})\bigg(a_{\bf k} e^{i{\bf k}\cdot{\bf r_n}}+a^\dagger_{\bf k} e^{-i{\bf k}\cdot{\bf r_n}}\bigg)C_{\bf n},
\label{hEM1}
\end{eqnarray}
Here we have defined $g({\bf k})=-i \sqrt{\frac{1}{2\hbar \omega_{\bf k} \epsilon_0\nu}}\omega_{0}\sum_\sigma {\bf e}_{{\bf k},\sigma}\cdot{\bf d}_{12}$, considering for simplicity that every atom has the same frequency $\omega_0$. We now assume also that the atoms are placed in a cubic lattice separated by a distance $d_0$ in each direction, so that ${\bf r}_n=d_0 {\bf n}=d_0(n_x,n_y,n_z)$, and the sums in ${\bf n}$ go from $n_\beta=1,\cdots, N_\beta$, with $N_\beta$ the number of atoms considered in the direction $\beta=x,y,z$. The emission of atoms in such a regular crystal-like structure have also been analyzed in \cite{porras2008}. The wave vector ${\bf k}$ takes the values $k_\beta=2\pi q_\beta/L_\beta$, with $L_\beta=h_0M_\beta$ defining the physical quantization region in the direction $\beta$, and $q_\beta=1,\cdots,M_\beta$. For simplicity, we shall define $M_\beta=M$, and $N_\beta=N$, and $d_0=h_0$.
Let us consider in addition that $g({\bf k})\approx g(k)\approx g(k_\textmd{eff})=g$, where $k_\textmd{eff}$ is the resonant wave-vector. As in the one dimensional case, the former is a good approximation when the phase $\exp(i\omega(k)\tau)$ varies much faster in ${\bf k}$ (both in module and angular dependency) than the coupling coefficient, so that
\begin{eqnarray}
\alpha_{nm}(\tau)&=&\gamma(\frac{h_0}{2\pi})^3\sum_{\sigma}\int d{\bf k}\frac{ |\hat{e}_{{\bf k},\sigma}\cdot \hat{u}_d|^{2}}{\omega(k)}e^{i{\bf k}\cdot{\bf r}_{nm}-i\omega(k)\tau}\cr
&\approx&\gamma\int d{\bf k} e^{i{\bf k}\cdot{\bf r}_{nm}-i\omega(k)\tau}
\label{corr3D}
\end{eqnarray}
where $\gamma=\hat{\gamma}(\frac{h_0}{2 \pi})^3 \sum_{\sigma}\frac{ |\hat{e}_{k_\textmd{eff},\sigma}\cdot \hat{u}_d|^{2}}{\omega(k_\textmd{eff})}$, with $\hat{\gamma}=\omega_{0}^2 d^2_{21}/(2\hbar\epsilon_0\nu)$, is a good approximation to the correlation function. Note that the quantization volume $\nu=h_0^3$.
In the following, we shall consider two different cases: an anisotropic dispersion relation, and an isotropic one.
\subsection{Anisotropic dispersion relations}
For an anisotropic dispersion relation, it is consider to assume a three dimensional Foruier transform $a_{\bf k}=\sum_{\bf m}e^{i{\bf k}\cdot {\bf r}_m}b_{\bf m}$, which leads to the simple form
\begin{eqnarray}
H=H_S+\sum^M_{\bf n,m=1}f_{\bf nm} b_{\bf n}^\dagger b_{\bf m}+g\sum_{\bf n} (b_{\bf n}^\dagger+b_{\bf n})C_{\bf n}.
\label{fintrans3D}
\end{eqnarray}
Here, the quantity
\begin{eqnarray}
f_{\bf nm}=\sum_{\bf k}\omega({\bf k})e^{-i{\bf k}\cdot{\bf r}_{nm}}
\end{eqnarray}
with ${\bf r}_{nm}={\bf r}_n-{\bf r}_m$. Notice that for anisotropic dispersion relations of the form $\omega({\bf k})=\omega(k_x)+\omega(k_y)+\omega(k_z)$, so that the transformed harmonic oscillators will conform a structure that for next neighbors interaction is just a cubic lattice (see this example in Fig. \ref{cubic2D}).
\begin{figure}[ht]
\centerline{\includegraphics[width=0.45\textwidth]{cubic2D.pdf}}
\caption{(Color online) A crystal of atoms and transformed modes conforming a cubic structure. It represents a mapping of the form (\ref{fintrans3D}) for an ensemble of atoms arranged in positions ${\bf r}_n={\bf n}h_0$, coupled to the radiation field within a $2D$ photonic crystal, which has dispersion relation $\omega({\bf k})=A+B_x\cos{(k_x h_0)}+B_y\cos{(k_y h_0)}$ \cite{devega2005} (the generalization to three dimensions is straightforward). The representation of the light-matter interaction problem here proposed may be useful to design and analyse schemes to control the flow of light and its absorption. \label{cubic2D}}
\end{figure}
\subsection{Isotropic dispersion relations}
For isotropic dispersion relations, if the atoms are in a one dimensional structure, the problem can still be mapped to a ladder-like structure. Let us write the interaction Hamiltonian (\ref{hEM1}) as
\begin{eqnarray}
H^\textmd{3D}_\textmd{int}=\frac{g}{\sqrt{M}} \sum_n \sum_q\int d\Omega_{\bf k}\bigg(a_{\bf k} e^{ik_q r_n}+h.c.\bigg)C_n.
\label{hEM1}
\end{eqnarray}
where we have assumed that the atoms are placed along the $z$ axis, and that the momentum modulus is discretized with an index $q$. In addition, we consider an isotropy correlation function $\omega({\bf k})=\omega(k)$, and
$g({\bf k})=g$.
Also, $g(k)=g$ is a good approximation when $\omega(k)$ gives rise, in the calculus of $\alpha_{nm}^\textmd{3D}(t)$, fo a phase that varies much more fast in $k$ than $g(k)$, which can be considered as a constant in comparison \cite{devega2005}.
Let us consider the transformation $a_{\bf k}=\int d\Omega_{r_p} \sum_{p=0,N-1}U_{\bf p}({\bf k}) b_{\bf r_p}$, with
\begin{eqnarray}
U_{\bf p}({\bf k})= \frac{1}{\sqrt{M}}\sum_{l=0}^\infty \sum_{m=-l}^l i^l Y_{lm}(\theta_p,\phi_p)Y_{l,m}^*(\theta_k,\phi_k)e^{-i k_q r_p},\cr
\end{eqnarray}
where $(r_n,\theta_n,\phi_n)$ and $(k,\theta_k,\phi_k)$ are ${\bf r}_n$ and ${\bf k}$ in spherical coordinates. Also, $Y_{lm}(\theta,\phi)$ are spherical harmonics. With this transformation, the interaction Hamiltonian (\ref{hEM1}) can be written as
\begin{eqnarray}
H^\textmd{3D}_\textmd{int}&=&\hat{g}\int d\Omega_{r_p} \sum_{p=0,M-1} \sum_q\int d\Omega_{\bf k}\bigg(\sum_{l=0}^\infty \sum_{m=-l}^l i^l \cr
&\times&Y_{lm}(\theta_p,\phi_p)Y_{l,m}^*(\theta_k,\phi_k)e^{-i k_q (r_p-r_n)}b_{{\bf r}_p}+h.c.\bigg)C_n.\cr
\label{hEM2}
\end{eqnarray}
where $\hat{g}=g/M$.
Interestingly, $Y_{00}(\theta,\phi)=\frac{1}{2\sqrt{\pi}}$ for any angle $\theta,\phi$. Thus, inserting $2\sqrt{\pi}Y_{00}(\theta_k,\phi_k)$ in the above expression we can solve the angular integral in $\Omega_k$,
\begin{eqnarray}
\int d\Omega_k Y_{0,0}(\theta_k,\phi_k)Y^*_{l,m}(\theta_k,\phi_k)=\delta_{l0}\delta_{m0},
\end{eqnarray}
with $\int d\Omega_k=\int_0^\pi d\theta\sin(\theta)\int_0^{2\pi} d\phi$, where we have used the property
\begin{eqnarray}
\int d\Omega_k Y_{l,m}(\theta_k,\phi_k)Y^*_{l',m'}(\theta_k,\phi_k)=\delta_{ll'}\delta_{mm'}.
\label{propgen}
\end{eqnarray}
Note that the property
\begin{eqnarray}
\sum_{lm} Y^*_{l,m}(\theta_p,\phi_p)Y_{l,m}(\theta_n,\phi_n)=\delta(\theta_p-\theta_n)\delta(\phi_p-\phi_n).
\end{eqnarray}
is also needed to show that the transformed operators $b_{{\bf r}_p}$ fulfill the proper conmutation relations, so that the transformation is canonical.
In addition, assuming that $|k|\equiv k_q=2\pi q/(Md_0)$, we find that $ \sum^M_{q=0} e^{i k_q (r_n- r_p)}=M\delta_{np}$. Hence, we find that
\begin{eqnarray}
H^\textmd{3D}_\textmd{int}&=&g\sum_n\bigg(B_{0,0,n}+B^\dagger_{0,0,n}\bigg)C_{n},
\label{hEM6sp}
\end{eqnarray}
where we have defined $B_{l,m,r_n}=\int d\Omega Y_{lm}(\theta,\phi) b_{r_n,\theta,\phi}$.
Note that because of the property (\ref{propgen}), this operators obey the usual bosonic commutation rules $[B_{l,m,n},B^\dagger_{l',m',n'}]=\delta_{ll'}\delta_{mm'}\delta_{nn'}$.
Let us now consider $H_B=\sum_{\bf k}\omega({\bf k})a^\dagger_{\bf k}a_{\bf k}$, and transform it as before
\begin{eqnarray}
H_B&=&\sum_{n,p,q}\int d\Omega_k\omega(k_q)\int d\Omega_p\int d\Omega_n U^*_{\bf n}({\bf k})U_{\bf p}({\bf k}) b^\dagger_{\bf r_n} \cr
&\times&b_{\bf r_p}.
\label{hBsp}
\end{eqnarray}
In detail, it can be written as
\begin{eqnarray}
H_B&=&\frac{1}{M}\sum_{n,p,q} \omega(k_q)\int d\Omega_p\int d\Omega_n \sum_{l,l'} \sum_{m,m'} (-i)^l(i^{l'})\cr
&\times& Y^*_{lm}(\theta_n,\phi_n)Y_{l'm'}(\theta_p,\phi_p)F_{ll'mm'}\cr
&\times&e^{i k_q (r_n-r_p)}b^\dagger_{\bf r_n} b_{\bf r_p}.
\label{hBsp2}
\end{eqnarray}
where $F_{ll'mm'}=\int d\Omega_k Y_{l,m}(\theta_k,\phi_k) Y_{l',m'}^*(\theta_k,\phi_k)=\delta_{ll'}\delta_{mm'}$ according to (\ref{propgen}). Hence, we find that
\begin{eqnarray}
H_B&=&\frac{1}{M}\sum_{l,m} \sum_{n,p}\sum_q\omega(k_q)e^{i k_q (r_n-r_p)}\int d\Omega_p \cr
&\times& Y^*_{lm}(\theta_n,\phi_n) b^\dagger_{\bf r_n} \int d\Omega_n Y_{lm}(\theta_p,\phi_p)b_{\bf r_p}.
\label{hBsp3a}
\end{eqnarray}
Or simply,
\begin{eqnarray}
H_B&=&\sum_{lm}\sum_{np} f_{np} B^{\dagger}_{l,m,r_n}B_{l,m,r_p}
\label{hBsp3}
\end{eqnarray}
where
\begin{eqnarray}
f_{np}=\frac{1}{M}\sum_q \omega(k_q) e^{i k_q (r_n-r_p)}.
\label{coefmap}
\end{eqnarray}
Indeed, from (\ref{hEM6sp}) and (\ref{hBsp3}), the only modes involved in the dynamics are the isotropic modes, $B_{0,0,r_n}$ and $B^\dagger_{0,0,r_n}$, with discrete positions $r_n$. This shows that for an isotropic dispersion relation, the full $3D$ problem can be mapped into a ladder-like structure of the form (\ref{fintrans}) as in the $1D$ case.
\subsection{Limit of independent environments}
\label{indep}
There is a limit in which the different atoms within the structure evolve as if each of them were interacting with its own environment. This limit is defined when the correlation function decays very fast with the inter-particle distance, such that $\alpha_{nm}(t)\approx\delta_{nm}\alpha_{nm}(t)$.
Note that this case is one of the most common in quantum optics, and corresponds to atoms emitting independently to each others. In this particular case, the atoms can be assumed to be arranged in any spatial structure and not necessarily a cubic lattice.
An alternative way to find this limit, can be based on the mapped structure. Indeed, the evolution time scale of the OQS can be estimated as $T_\textmd{diss}\approx 1/\Gamma$, where $\Gamma\approx\textmd{Re}[\int_0^\infty ds\alpha_{nn}(s)]$. Then, the distance that an excitation go along one direction through the chain during that time is $L_\textmd{diss}\approx vT_\textmd{diss}$, where $v=d_0f$ is the velocity of the excitation within the chain, with $f$ the average hopping rate between sites. Hence, in the limit where $L_\textmd{diss}\ll d_0P$, or $f/\Gamma\ll P$, the excitation will not have time to travel to adjacent atoms within the structure during the dissipation time, and therefore each atom within the ensemble will be coupled to its own environment.
In the above discussed limit, the effects of the environment are integrally encoded in the spectral function $J(\omega)=g^{2}(k(\omega))\rho_\textmd{DOS}(\omega)$, where $\rho_\textmd{DOS}(\omega)=A(k(\omega))=|\frac{d \omega(k)}{dk}|^{-1}_{k=k(\omega)}$, as $A(k)=|\frac{d \omega(k)}{dk}|^{-1}$ \cite{chin2010}. Thus, different pairs of $g(k)$ and $\omega(k)$ can lead to the same $J(\omega)$, and we can consider $g(k)=g$ in particular, together with a new $\hat{\omega}(k)$ such that $|\frac{d \hat{\omega}(k)}{dk}|^{-1}_{k=k(\omega)}=J(\omega)$, where now $k(\omega)$ is the inverse of $\hat{\omega}(k)$ (See Appendix B for an example).
\section{Link to coupled cavity QED}
When considering $H_S=\sum_n\omega_n\sigma_n^\dagger\sigma_n$, the Hamiltonian (\ref{fintrans}) has the form of that of $M$ cavities with a resonant mode $b_n$ and energy $f_n=f_{nn}$, each cavity containing a single atom of frequency $\omega_n$ coupled to the cavity mode with strength $g$. Then, the second term of (\ref{fintrans}) for $f_{nm}$ with $n\neq m$, can be seen as a coupling between the different cavities, where each cavity site is described by the well-known Jaynes-Cummings model.
The non-linearities in this system lead to an on-site repulsion, which combined with the hopping term between the cavity modes $b_n$ leads to a Bose-Hubbard-like dynamics where quantum phase transitions of light can be described in the ground state between a Mott-like phase (i.e. the photon blockade regime) and a superfluid phase
\cite{greentree2006,hartmann2006,angelakis2007,schiro2012}.
In addition, strong signatures of photon blockade have also been observed in the non-equilibrium dynamics of weakly coupled cavity arrays \cite{nissen2012}.
Thus, in the transformed system, we shall expect similar ground state and dynamics than in coupled cavity arrays. The main difference is that in the transformed system, the atoms may be directly coupled with each others through resonant dipole-dipole interactions, while in the coupled cavity case this might be more difficult to achieve. Also, as discussed above, for some particular environments the modes may be connected beyond next neighbors.
For the case of two dimensional environments, the transformed system may also give rise to a coupled cavity array of the form of Fig. (\ref{cubic2D}) provided that the dispersion relation can be written as $\omega({\bf k})=\omega(k_x)+\omega(k_y)$. Then, the coupling $f_{{\bf n m}}=f_{n_x,m_x}\delta_{n_y,m_y}+f_{n_y,m_y}\delta_{n_x,m_x}$, i.e. it will only connect sites within the $x$ or $y$ directions, but never across the diagonal. Note that a similar reasoning can be followed for the case of three dimensional environments.
\section{Application: Atoms within a 1D photonic crystal}
We shall consider the dynamics of atoms or quantum dots embedded in a $1D$ photonic crystal structure \cite{yablonovitch1987,john1987}, considering that the other two directions are non-dispersive. This problem can be described with the Hamiltonian (\ref{h1}), considering a dispersion relation of the form $\omega(k)=A+B\cos((k-k_0)h_0)$ \cite{devega2005}, where $k_0=\pi/h_0$, $h_0$ is the linear size of the unit cell of a cubic lattice, and $k$ runs from $-k_0$ to $k_0$ within the first Brillouin zone (see Fig. (\ref{dispersion})).
PCs typically consist of a low-dielectric-constant network, inserted in a high-dielectric-constant backbone in a periodic structure. There is a large variety of photonic crystals giving rise to a complete one, two or three dimensional band gap \cite{pcbook}.
Of special interest is the analysis and control of the spontaneous emission, which can be dramatically modified by the presence of such band-gap dispersion relation \cite{john1994,florescu2001,devega2005,giraldi2011}. This may have important applications ranging from miniature lasers and light-emitting diodes, to single-photon sources for quantum information, and to solar energy harvesting \cite{lodahl2004}. Particularly, experimental progress in the control of spontaneous emission by manipulating optical cavity modes and quantum dots within photonic crystals have demonstrated that the spontaneous emission from light emitters embedded in photonic crystals can be suppressed by the so-called photonic bandgap, whereas the emission efficiency in the direction where optical modes exist can be enhanced \cite{noda2007,thompson2013}. Partial dissipation is also observed in Fano-Anderson-like models, where one or more quantum emitters couples to a finite band of modes \cite{gaveau1995}. Recent proposals \cite{hung2013,goban2013} explore the atom-atom interactions that may be produced in these materials, and which are mediated by a strong light-matter interaction. Hence, being able to analyze such effects with analytic tools that allow to explore regimes beyond the Markov and weak coupling approximations may be of extreme importance to understand experiments and lead to further developments.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.4\textwidth]{DispersionPC3.pdf}}
\caption{(Color online) Dispersion relation of the radiation field within a photonic crystal. It shows the relationship between the frequency of the photons and the different values of the wave vector within the first Brillouin zone (ranging from $[-\pi/h_0,\pi/h_0]$). The dispersion relation is not defined for frequency values within the range from $0$ and $A-B$, and larger than $A+B$. These values correspond to a gap in the photonic density of states. The orange oval denotes to the approximate parameter region where $\delta/B\ll 1$, so that an effective mass approximation (EMA) can be considered (see discussion). \label{dispersion}}
\end{figure}
The correlation function corresponding to (\ref{h1}) can be written as (\ref{correl1D}) or, considering that the all atomic frequencies are $\omega_0$,
\begin{eqnarray}
\alpha_{nm}(\tau)&=&\hat{\gamma}(\frac{h_0}{2\pi})\sum_{\sigma}\int_{1BZ}dk\frac{ |\hat{e}_{{\bf k},\sigma}\cdot \hat{u}_d|^{2}}{\omega(k)}e^{ikr_{nm}-i\omega(k)\tau}\cr
&=&\gamma\int_{-\frac{\pi}{h_0}}^{\frac{\pi}{h_0}}dk e^{ikr_{nm}-i\omega(k)\tau},
\label{generalcorr5}
\end{eqnarray}
where we have used an approximation similar to (\ref{corr3D}), and defined again $\gamma=\hat{\gamma}(\frac{h_0}{2 \pi}) \sum_{\sigma}\frac{ |\hat{e}_{k_\textmd{eff},\sigma}\cdot \hat{u}_d|^{2}}{\omega(k_\textmd{eff})}$. Here, $\alpha_{nn}(t)=\gamma e^{-iAt}J_0(Bt)$, with $J_0(t)$ the zeroth order Bessel function.
The transformed Hamiltonian takes the form (\ref{fintrans}), with
$f_{nn}=\sum_{q=0}^{M-1} e^{-ik_qr_n}e^{ik_qr_n} \omega(k)=A$, and $f_{n,m}=\sum_{q=0}^{M-1} e^{-ik_qr_n}e^{ik_qr_m}\omega(k)=\delta_{m,n\pm 1}\frac{B}{2}$.
The resulting Hamiltonian is
\begin{eqnarray}
H_{\textit PC}&=&H_S+A\sum_n b_n^\dagger b_n+\frac{B}{2}\sum_{n=1,\cdots M}\bigg(b_n^\dagger b_{n+1}\cr
&+&b_{n+1}^\dagger b_{n}\bigg)+g\sum_{n=1,\cdots M} C_n (b_n+b_n^\dagger),
\label{mappedPC}
\end{eqnarray}
with $C_n=\sigma_n+\sigma_n^+$, and
\begin{eqnarray}
H_S=\sum_{j}\omega_0\sigma_j^+\sigma_j-\sum_{\langle jl\rangle}J\sigma_j^+\sigma_l.
\end{eqnarray}
For the sake of simplicity in the calculations, we shall consider the rotating wave approximation in the former Hamiltonian, so that the terms $b_n^\dagger \sigma^+$ and $b_n \sigma_n$, that simultaneously creates (and annihilates) one photon and one excitation in the atomic lattice are discarded.
\subsection{Units of the problem}
\label{units}
A brief comment is here in order regarding the units of the problem. The parameters that characterize impurities in a photonic crystal with a gap in the optical region \cite{florescu2001,devega2005} are $\omega_{0}\sim 10^{15}$ Hz, $d_{21}\sim 10^{-29}$Cm ($10^{-28}$Cm for quantum dots), $h_0\sim 10^{-6}-10^{-7}$m, and $A\sim 10^{15}$Hz. With these values, a realistic strength of the coupling constant appearing in the Hamiltonian is around $g\approx g_{n}(k_0)=-i \sqrt{\frac{1}{2\hbar \omega(k) \epsilon_0\nu}}\omega_0d_{12}$ of the order of GHz-THz. Note that this quantity shall not be confused with the decaying rate, which is given by $\Gamma\approx \textmd{Re}[\int_0^\infty ds\alpha_{nn}(s)]$, and therefore is related to the constant $\gamma$ in (\ref{generalcorr5}).
The correlation function, when appearing in the evolution equations in interaction image with respect to the system Hamiltonian can be re-written as $\alpha^\textmd{int}_{nm}(t)= \gamma\int_{-\frac{\pi}{h_0}}^{\frac{\pi}{h_0}}dk e^{ikr_{nm}+i(\Delta-B\cos((k-k_0)h_0)t)}$, where $\Delta=\omega_0-A$. Hence, the atomic dynamics will only depend on the different values of $\Delta$, $B$ and $g$. Here, we will chose an energy scale $\xi$, and assume that all the quantities are re-normalized in this scale, i.e. $\tilde{\Delta}=\delta/\xi$, $\tilde{J}=J/\xi$, $\tilde{B}=B/\xi$, $\tilde{g}=g/\xi$, and $\tilde{t}=t\xi$, although in the following the tilde is omitted for simplicity in the notation. The choice $\xi$ between the range of GHz-THz will lead us to typical values for photonic crystal coupling strengths, what for the chosen parameters will force the band width $B$ to be within a similar range. The choice of other scaling parameters may correspond to other applications different from photonic crystals, like for instance atoms in optical lattices \cite{devega2008,navarrete2010}.
In the limit where $(k-k_0)h_0\ll 1$ (i.e $k\ll 2/ h_0$), we may expand the dispersion relation $\omega(k)=A+B\cos(k-k_0)$ as $\omega_k\approx\omega_c+B/2(k-k_0)^2$ \cite{devega2005}. This last expression for the dispersion relation corresponds to the one obtained with the so-called effective mass approximation \cite{john1994,florescu2001}, and correspond in frequencies to the choice $\delta/B\ll 1$ (See Fig.(\ref{dispersion})).
Hence, in the effective mass approximation, the correlation function of the environment appearing in the atomic evolution equations in interaction image with respect to the system, is written as $\alpha^\textmd{int}_{nm}(t)\approx \gamma\int_{-\frac{\pi}{h_0}}^{\frac{\pi}{h_0}}dk e^{ikr_{nm}-i(\delta+B/2(k-k_0)^2)t}$, where $\delta=\omega_0-\omega_c$. In this case, the atomic dynamics only depend on the different values of $\delta$ (see also \cite{john1994,florescu2001} for more details), the band width $B$, and the coupling strength $g$.
\subsection{Comparison to the master equation}
Let us consider the solution of this problem according to a master equation. This equation describes the evolution of the reduced density operator of the atoms, which is obtained by tracing out all the environment degrees of freedom as $\rho_s(t)=\textmd{Tr}_B[\rho_\textmd{tot}(t)]$. Up to second order in the coupling parameter between system and environment, $g$, the master equation corresponding to (\ref{h1}) is given by \cite{breuerbook}
\begin{eqnarray}
\frac{d\rho_s (t)}{dt}&=&-i[H_S (t),\rho_s (t) ]\cr
&+&\int_0^t d\tau \sum_{lj} \alpha_{lj}(t-\tau)[L_j(\tau-t)\rho_s (t) ,L_l^{\dagger}]\nonumber\\
&+&\int_0^t d\tau \sum_{lj} \alpha^{*}_{lj} (t-\tau)[L_l,\rho_s (t) L_j^{\dagger}(\tau-t)].
\label{icc20}
\end{eqnarray}
with $L_j(t)=e^{iH_S t}L_je^{-iH_S t}$, and $\alpha_{lj} (t-\tau)= g^2\sum_k e^{i k r_{lj}-i\omega_{\bf k} (t-\tau)}$, where $r_{lj}=d_0(l-j)$. To derive this equation, the so-called Born approximation has been assumed. Hence, the correlations between the system and the environment have been neglected, so that $\rho_\textmd{tot}(t)=\rho_s(t)\otimes\rho_B$, where $\rho_B$ is the environment density operator considered always in its equilibrium state.
We now compare in Fig. (\ref{chainPC0}) the population dynamics of $N$ atoms given by the master equation (\ref{icc20}), and by a Schr{\"o}dinger equation for the mapped Hamiltonian $H_{\textit PC}$.
This simple example shows the power of the proposed scheme. While the master equation clearly fails after a few time steps giving rise, in some cases, to non-physical results (having a reduced density matrix $\rho_S$ with negative eigenvalues), the Schr{\"o}dinger equation for $H_{\textit PC}$ gives the correct evolution. The number of oscillators needed is of the order of the time scale to be reached, i.e. $M=100$ oscillators.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.45\textwidth]{compara3.pdf}}
\caption{(Color online) Evolution of the population of two atoms (black and red curves respectively), for $B=0.5$, $A=1$, and $g=0.05$ (upper plot) and $g=0.1$ (lower plot), considering $\omega_0=0.3$, $J=0.$, and $|\psi_0\rangle=|1\rangle|0\rangle$. Solid and dashed curves correspond respectively to the result obtained using $H_{\textit PC}$ or a master equation up to $g^2$ \cite{breuerbook} derived for (\ref{h1}). Black, grey (with divergent dashed curve in the lower panel) and orange curves correspond to $B=0.5,0.8,1$ respectively. Dotted lines in the plot below represent the sum of negative eigenvalues of the reduced density operator $\rho_S(t)$ obtained from the master equation. For $B=0.8$ the quantity grows very large, indicating the inaccuracy of the result. See units in Sect. (\ref{units}). \label{chainPC0}}
\end{figure}
We have seen that this model shows a very rich dynamics that the mapping unveils well beyond the weak coupling approximation. In addition, it provides a tool to analyze the formation of cavities within photonic crystal structures, particularly within the gap region. As it can already be seen in Fig (\ref{chainPC0}), for certain parameter regimes the atomic population does not fully decay, which is connected to the fact that the atom is interacting with just a few environment modes.
The mapping shows directly that these few environment modes are just the ones within the chain which are located in the neighborhood of the atom or emitter. We will discuss these ideas in the following for two different cases, $N=2$ and $N=15$ atoms and a single excitation in the system.
\subsection{Formation of cavities}
The appearance of a non-zero steady state population depends highly on whether the atomic frequency is within the band or the gap (see Fig. (\ref{dispersion})). This can be seen in Fig. (\ref{chain3D}), that presents a density plot of the time average of the population $P_T(t)=\frac{1}{t}\sum_{j=1,N}\int_0^t ds\langle\sigma_j^+(t)\sigma_j(t)\rangle$ at $t=300$ with respect to $B$ and $\omega_0$. When varying the hopping rate $B$, a crossover is observed between the regime where the atomic population does not vanish in the long time limit, and a regime where full relaxation is observed. Indeed, a similar crossover was observed in the region of small $\delta=\omega_0-\omega_c$ \cite{john1994},
when analyzing the problem within the effective mass approximation. Also, as seeing in Fig. (\ref{chain3D}), at longer times the contrast between the lighter regions (with non-vanishing atomic population), and the darker regions (with vanishing atomic population) is expected to become stronger.
In other words, the black region in the upper panel corresponds to parameters where the atomic population is slowly decaying to zero. To see this, the lower panel shows the time evolution of $P_T(t)$ at even longer times ($t=800$) than the time at which the upper density plot is represented. Such time evolution is displayed for the parameters corresponding to the points marked in circles in the upper figure. It is observed that for $B=0.6$ the values of $P_T(t)$ at $t=800$ (of the order of $0.1$ and $0.5$) are smaller than the ones for $t=300$ ($0.2$ and $0.1$ respectively), showing a slow decaying.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.45\textwidth]{new3Db3.pdf}}
\caption{(Color online) Upper plot: Density plot of the time average of the total population $P_T(t)$ at $t=300$ with $\omega_0$ and $B$, with $A=1.$, $g=0.1$ and $J=0$. The lower panel represents the time evolution of $P_T(t)$ at even longer times ($t=800$) for the parameters corresponding to the points marked in the upper figure, i.e. $B=0.6$ and $\omega_0=0,0.4,0.5,1$ (solid black, dashed red, dotted blue and dot-dashed orange respectively). See units in Sect. (\ref{units}).
\label{chain3D}}
\end{figure}
The incomplete relaxation occurring for atomic frequencies within the gap occurs because of the presence of polaritons (known as photon-atom bound states in the photonic crystal literature \cite{john1994,devega2005}). Due to these polaritons, part of the energy initially in the atomic system stays trapped nearby the atoms, and therefore it is never irreversibly lost. The presence of these polaritons, which represent highly correlated atom-photon states, explains partially the failure of the master equation (\ref{icc20}), that as mentioned above is based on the Born approximation and thus neglect any system-environment correlation. Atom-photon bound states are known to lead to non-Markovian dynamics \cite{tong2010}, and such non-Markovianity can be even qualified by the recently derived non-Markovianity measures \cite{breuer2009,rivas2009,rivas2014}. This is done for instance in \cite{tufarelli2014} for the case of an atom radiating in a one-dimensional photonic crystal waveguide in the presence of a mirror. However, it is important to note that even if such bound states are not formed, non-Markovianity can still be significant. This occurs for instance in an atom coupled to a cavity mode, a system where bound states are not formed but which still presents a strong non-Markovianity \cite{laine2010} (see also \cite{lambropoulos2000}). In our case, at the parameter regimes where the system has no atom-photon bound states, the excitations will flow away through the harmonic oscillator chain so that the OQS will eventually relax to its ground state. However, as noted above some back-flow of information (leading to a temporal increase of the OQS’ coherences) may occur at the environment correlation time scales, and thus non-Markovian effects may still be present.
The mapped Hamiltonian (\ref{mappedPC}) allows to see clearly the presence of such polaritons in the system. Considering a single excitation, the system basis can be written as a set of atomic-type of states ($|\psi^\textmd{at}_1\rangle=|1,0\rangle|vac\rangle$ and $|\psi^\textmd{at}_2\rangle=|0,1\rangle|vac\rangle$ for $N=2)$, where the excitation is in the atomic degrees of freedom, and environment type of states ($|\psi^\textmd{env}_j\rangle=b_j^\dagger|0,0,\cdots,0\rangle$, for $j=N+1,\cdots,M$), where the excitation is contained within one of the chain modes. An exact diagonalization of the Hamiltonian in such basis gives rise to four eigenvectors $|P_j\rangle$ that are combination of states $|\psi^\textmd{at}_n\rangle$ where the excitation is in the atoms, and states $|\psi^\textmd{env}_j\rangle$ where the excitation is in near-neighbor modes of the atoms. Here we consider these eigenvectors as polaritons, in the sense that they are conformed by an atomic part and a photonic part. In addition, because they are both eigenstates of $H$ and have no overlap with sites beyond near-neighbors, the subspace they span is what is known in the literature as an invariant subspace \cite{caruso2009}. In such subspaces the excitations are trapped and never flow away to other states external to the subspace. Naturally, if the system is initially prepared in one of the invariant states, then there will be no dynamics. Similarly, the fraction of the initial state corresponding to this invariant states will not vary with the dynamics. Indeed, the initial state $|\Psi_0\rangle=|\psi_1^\textmd{at}\rangle$ (corresponding to the first atom initially excited) can be proyected in the basis of polaritonic $|P_j\rangle$ and non-polaritonic or non-invariant eigenvectors $|\phi_j\rangle$, what leads to $|\Psi_0\rangle=\sum_j a_j |P_j\rangle+\sum_j b_j|\phi_j\rangle$, where $a_j=\langle P_j|\Psi_0\rangle$ and $b_j=\langle\phi_j|\Psi_0\rangle$. Then, because of the non-degeneracy of the Hamiltonian spectra, the amount of population trapped in the invariant subspace can be calculated just as $P_{\textmd{pol}}=\sum_j |a_j|^2$. Since polaritons combine atomic and photonic degrees of freedom, this quantity is an upper bound to the final state population within the atomic system.
Fig. (\ref{polar1}) represents $P_{\textmd{pol}}$ with zero hopping rate (upper plot), and finite hopping rate (lower plot). The upper panel presents a very similar profile to the upper panel of (\ref{chain3D}), which reflects the atomic population at long times ($t=300$). The only difference is that in the upper plot of (\ref{polar1}), the population within the band (i.e. for frequencies $A-B< \omega_0< A+B$) vanishes completely, whereas in the plot representing the atomic population the population has not jet decayed completely at $t=300$. A further difference appears to be the fact that the polariton analysis predicts a smaller steady state population, of the order of $0.5$, in the upper gap (i.e. for frequencies above $A+B$) than in the lower gap (i.e. for frequencies below $A-B$), whereas the upper panel of (\ref{chain3D}) shows an equal solution for both gaps. Hence, it can be concluded that the upper gap suffers some decaying, that is nevertheless very slow and cannot be captured at the time scale in which (\ref{chain3D}) is plotted.
The lower panel in in Fig. (\ref{polar1}) gives the polariton population when considering $J=0.25$. Some of the features within this plot can be qualitatively explained. For instance, the fact that the black region is displaced with respect to the black region in the upper panel, can be explained because the eigen-energies of the new system Hamiltonian, $\hat{H}_S=(\omega_S-\Delta_\textmd{LS})(|0,1\rangle\langle 0,1|+|1,0\rangle\langle 1,0|)+\omega_g|0,0\rangle\langle 0,0|-J\sum_j (|1,0\rangle\langle 0,1|+|0,1\rangle\langle 1,0|)$, and therefore the relevant energy transitions are no longer $\omega_0$ and $0$ (or $\omega_s$ and $\omega_g$ without a re-normalization such that $\omega_0=\omega_s-\omega_g$). In the effective system Hamiltonian, the action of the environment can be included approximately as a Lamb shift, $\Delta_\textmd{LS}={\mathcal Im}[\Gamma_0]$, with $\Gamma_0=\int_0^\infty d\tau\alpha_{jj}(\tau)$ the so-called dissipation rate, and $\alpha_{jl}(t)$ given by (\ref{generalcorr5}). Indeed, when diagonalizing $\hat{H}_S$, we find the eigenvectors $|\phi_0\rangle=|0,0\rangle$, $|\phi_1\rangle=\frac{1}{\sqrt{2}}(|0,1\rangle-|1,0\rangle)$, and $|\phi_2\rangle=\frac{1}{\sqrt{2}}(|0,1\rangle+|1,0\rangle)$ corresponding to the eigen-energies $E_0=\omega_g$, $E_1=\omega_S-\Delta_\textmd{LS}-J$ and $E_2=\omega_S-\Delta_\textmd{LS}+J$ respectively. Hence, two different transitions of the OQS shall be considered, $\Delta_1=E_1-E_0=\omega_0-J$, and $\Delta_2=E_2-E_0=\omega_0+J$ (here we have discarded the Lamb shift, that can be neglected for sufficiently small couplings). When both transitions lie within the lower gap, i.e. $\Delta_1<A-B$, $\Delta_2<A-B$, we are in observe full preservation of the polariton population. For the parameters in Fig. (\ref{polar1}), this corresponds to the case when $\omega_0<0.25$ (white region in the lower panel). However, when $\Delta_1$ is within the gap, but $\Delta_2$ is within the band, i.e. in the region where $A-B-J<\omega_0<A-B+J$, only half of the population is lost, and therefore $P_{\textmd{pol}}=0.5$. The quantities at the two sides of the inequality mark the boundaries of the orange region within the lower panel of Fig. (\ref{polar1}), in our case given by $0.25<\omega_0<0.75$.
In addition, Fig. (\ref{polar2}) shows the amount of population trapped in the form of polariton depending on $J$ and for different values of the atomic frequency. In detail, the curves correspond to values of $\omega_0$ ranging from $0.1$ to $0.5$ (and from $0.6$ to $1$ in the inset) with intervals of $\Delta\omega_0=0.1$. It can be observed that there is a certain value of $J$, which depends on $\omega_0$, up to which the trapped population remains equal to $0.5$. Indeed, following a similar analysis as before, it can be concluded that the full width at half maximum of each curve is given by the point in energy where the transition $\Delta_2$ enters into the band, while $\Delta_1$ remains in the gap. This corresponds for instance to $J=A-B-\omega_0=0.4$ for $\omega_0=0.1$ or $J=0.1$ for $\omega_0=0.4$, and also explains why the different curves are displaced by $\Delta\omega_0$. A similar analysis can be made to analyse the insert within the figure, which shows the permanence of polaritons for atomic frequencies within the band. The proportion of atomic component in the polariton varies for each particular values of $\omega_0$ and $J$. This can be seen from the dotted curves, which represent the total atomic population $P_T(t)$ at $t=500$. For certain values of $J$ this atomic population is still higher than the population within the polariton, which reflects the fact that the population has not completely relaxed to its steady state value (not shown here).
\begin{figure}[ht]
\centerline{\includegraphics[width=0.45\textwidth]{New3DJ0J0_25.pdf}}
\caption{(Color online) Density plot that represents the total population trapped in the polariton state $P_{\textmd{pol}}$ with respect to $\omega_0$ and $B$, when considering $g=0.1$, $A=1$, $N=2$ and the initial state $|\Psi_0\rangle=|1\rangle|0\rangle|{\textmd vac}\rangle$. The upper plot corresponds to the case when $J=0$, and the lower plot represents the case $J=0.25$. See units in Sect. (\ref{units}).
\label{polar1}}
\end{figure}
\begin{figure}[ht]
\centerline{\includegraphics[width=0.45\textwidth]{PolaritonvariosJ2.pdf}}
\caption{(Color online) Solid curves represent the total population in trapped in the form of polaritons $P_{\textmd{pol}}$, for different hopping rates $J$ and different atomic frequencies. The different curves correspond to values of $\omega_0$ ranging from $0.1$ (curve in the right extreme) to $0.5$ (curve in the left extreme) in an interval of $\Delta\omega_0=0.1$. These curves represent therefore values within the gap, while the inset represents $\omega_0$ within the band, ranging from $0.6$ (left extreme) to $1$ (right extreme). The dotted lines represent the total atomic population $P_T(t)$ at $t=500$ under the same conditions. In all curves we have considered $N=2$ atoms, $A=1$, $B=0.5$, and $g=0.1$. The initial condition is considered $|\Psi_0\rangle=|1\rangle|0\rangle|{\textmd vac}\rangle$, where $|{\textmd vac}\rangle$ is the vacuum state for the environment. See units in Sect. (\ref{units}).\label{polar2}}
\end{figure}
Let us now analyse further the formation of cavities inside the photonic crystal. To this order, we consider in figure (\ref{trapped}) a histogram of the atomic and photonic populations in the long time limit for two different situations: when the atomic frequency is $\omega_0=0.1$, i.e. deep inside the gap (left plots), and when it is just in the band-gap edge, $\omega_0=B=0.5$ (right plots). The two cases present very different results. While for for $\omega_0=0.1$ the population remains localized nearby the initially excited atom, for the resonant case $\omega_0=B=0.5$, the population spreads along the whole atomic sample, and along more modes within the chain. This result confirms numerically the formal discussion in \cite{john1990}, showing that indeed the localization length $\xi$ of a photon grows larger and eventually diverges near the band-gap edge $\xi\sim 1/(\sqrt{\omega_c|\omega_c-\omega_0|})$. Here, it can be seen that deep inside the gap, the excitation remains localized nearby the original location (see also \cite{devega2008c}), which effectively corresponds to the formation of a cavity where the cavity mode is just the transformed oscillator coupled to the atom. A similar result is observed for both cases when one and two excitations are initially present in the atomic lattice.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.5\textwidth]{histotunned.pdf}}
\caption{(Color online) Histograms representing the time-averaged atomic (dotted orange) and photonic (plain violet) populations within the long time limit. The left and right plots correspond to $\omega_0=0.1$ and $0.5$ respectively. The upper panels correspond to the case where $N=2$ atoms, considering one atom initially excited (and two for the insets). The lower panels represent the case of $N=15$ atoms, and a single initial excitation in the system. Other parameters considered are $M=50$, $B=0.5$ and $A=1$ as in previous cases. \label{trapped}}
\end{figure}
\section{Conclusions and perspectives}
The paper proposes a method to map a many body OQS conformed by a regular particle array into that of two coupled lattices, corresponding to the OQS and its environment respectively. The atoms within the OQS are then directly coupled to \textit{the most relevant modes of the environment}, which in principle may allow to solve the problem with a variety of techniques alternative to the master equation.
The mapping is particularly simple for atoms in contact with the electromagnetic field within a photonic crystal, that lead to interactions within the environment modes that extend only to next-neighbours.
Generally speaking, in the case of bosonic environments, the mapping may lead to a system similar to the Jahn-Teller \cite{greentree2006,porras2012,nevado2012} and the Jaynes-Cummings for coupled cavities.
Hence, re-expressing an OQS Hamiltonian as (\ref{fintrans}) allows to use both numerical and analytical tools alternative to the ones traditionally considered in the analysis of OQS, that can be helpful to analyse the system beyond the usual weak coupling and Markovian regimes.
These ideas are illustrated in this paper by analysing the dynamics of atoms coupled to photonic crystals with a one dimensional gap. For this particular system, the result given by the master equation approach is compared to the result given by solving the Schr{\"o}dinger equation for the mapped system. Also, because the mapping allows to truncate to the most relevant modes of the environment, an exact diagonalization of the total Hamiltonian is carried out, which unveils the existence of highly correlated atom-photon states in the system when the atomic frequencies are within the gap, and a strong dependence of such highly correlated states on the presence of inter-atomic dipole-dipole interactions.
Also, it is noted that the transformed system is very similar to the starting setup from which collision-model-based non-Markovian master equations are derived \cite{giovannetti2012,ciccarello2013}. These models give rise to master equations that preserve highly desirable mathematical properties such as complete positivity, but have the drawback that they are not easily derivable from microscopic models. For instance, the quantities involved in the collisional model approach cannot in general be expressed in terms of the environment spectral density. Also, collisional models cannot be easily extended to deal with a many body OQS. Hence, the chain-mapping approach here proposed could be used as a starting point to derive a new collision-model-based master equation that on the one hand preserves complete positivity, and on the other hand can be extended to the many particle case and be directly related to a microscopic derivation that departs from first principles, i.e. from the total Hamiltonian of the system and its environment.
Being able to analyze the dynamics of many body OQS may bring new insight to a wide variety of problems, ranging from quantum optics (e.g. analysis of the dynamics of impurities in structured environments such as photonic crystals \cite{john1995,john1997}) and solid state physics (e.g. analysis of the dynamics of superconducting qubits \cite{you2005,you2012} and of strongly correlated systems with dissipation \cite{calzadilla2006,prosen2008,daley2009,schwanger2013,cai2013}), to quantum biophysics (e.g. study of the energy transport within photosyntetic complexes \cite{chin2013,rey2012}). In addition, understanding the dissipative dynamics beyond the Markov approximation is of primary importance to further develop the concepts of dissipative quantum computation and state preparation developed in \cite{verstraete2009,diehl2008,diehl2010}, and experimentally realized in \cite{subasi2012}.
Finally, the appealing form of Hamiltonian (\ref{fintrans}) suggests that the dynamics of OQS can be simulated with optical lattices in a spirit similar to the proposals \cite{recati2005,devega2008}, provided that the interaction strengths $f_{nm}$ corresponding to a particular environment are implemented. An alternative implementation of this system, consisting in a regular lattice of atoms connected to an environment, are Coulomb crystals of trapped ions \cite{porras2008,mavadia2013}.
\textit{Acknowledgements}
The author gratefully acknowledges D. Alonso, M.C. Ba{\~n}uls, C. Busser, A. Gonzalez-Tudela, A. Perez, C. Zi and U. Schollw{\"o}ck for interesting discussions, and D. Alonso, J.I. Cirac, A. Ekert
and U. Schollw{\"o}ck for encouragement and support. This project was financially supported by the Nanosystems Initiative Munich (NIM) (project No. 862050-2) and partially by the Spanish MICINN (Grant No. FIS2010-19998).
\section*{Appendix A: System evolution equations and their dependency on the environment correlation function}
Let us now follow the derivation in \cite{alonso2007} to show that for initially thermal states, the only relevant quantity to describe the coupling with the environment is the so-called correlation function. To see this, we consider the simple case of an atom coupled to an environment with a Hamiltonian of the form (\ref{hEM1}). The idea is to consider the Heisenberg equation for a system operator, $A$, and re-express it in such a way that the environmental operators $a_{\bf k} (0)$ are placed on the right hand side of the terms, while the $a^\dagger_{\bf k} (0)$ appear in the left hand side. Thus, when computing $\langle A(t)\rangle=\textmd{Tr}[A(t)\rho_0]$, with $\rho_0=\rho_S\otimes\rho_T$, and $\rho_T$ a thermal state for the environment, these terms vanish.
The Heisenberg evolution equation of $A(t)={\mathcal U}^{-1}(t,0)A{\mathcal U}(t,0)$, with ${\mathcal U}(t,0)$ the evolution operator with the total Hamiltonian H, can be written as
\begin{eqnarray}
&&\frac{dA(t_1 )}{dt_1}=i{\mathcal U}^{-1}(t_1 ,0 )[H_{tot},A]{\mathcal U}(t_1 ,0 )\cr
&=&-i [H_S (t_1 ),A(t_1 )]+i\sum_{\bf k} g({\bf k}) \large( a_{\bf k}^{\dagger} (t_1,0 )[L(t_1 ),A(t_1 )]\cr
&+&[L^{\dagger}(t_1 ),A(t_1 )]a_{\bf k} (t_1 ,0 ) \large),
\label{eq39}
\end{eqnarray}
We can replace in (\ref{eq39}) the formal solution of the evolution equation of the environmental operators, $da_{\bf k} (t_1 ,0)/dt_1 =i[H_{tot}(t_1 ),a_{\bf k} (t_1,0 )]=-i\omega_{\bf k} a_{\bf k} (t_1,0 )-ig({\bf k}) L(t_1 )$,
\begin{eqnarray}
a_{\bf k} (t_1 ,0)&=&e^{-i\omega_{\bf k} t_1 }a_{\bf k}(0,0)-i g({\bf k}) \int^{t_1}_0 d\tau e^{-i\omega_{\bf k} (t_1-\tau)}\cr
&\times&L(\tau).
\label{eq40}
\end{eqnarray}
The single evolution equation (\ref{eq39}) becomes as follows,
\begin{eqnarray}
&&\frac{dA(t_1 )}{dt_1}=i [H_S (t_1 ),A(t_1 )]-\nu^{\dagger}(t_1 )[L(t_1 ),A(t_1 )]\nonumber \\
&+&\int_0^{t_1 }d\tau \alpha^*(t_1 -\tau) L^{\dagger}(\tau)[A(t_1 ),L(t_1 )]+[L^{\dagger}(t_1 ),A(t_1 )]\cr
&\times&\nu(t_1 )+\int_0^{t_1 }d\tau \alpha(t_1 -\tau)[L^{\dagger}(t_1 ),A(t_1 )]L(\tau),
\label{eq41}
\end{eqnarray}
where we have defined the environment correlation function as
\begin{equation}
\alpha(t-\tau)=\sum_{\bf k} |g({\bf k})|^2 e^{-i \omega_{\bf k} (t-\tau)}.
\label{chapuno325}
\end{equation}
In the last expression, we have also defined the bath operators
\begin{eqnarray}
\nu^{\dagger}(t_1 )=-i\sum_{\bf k} g({\bf k}) a_{\bf k}^{\dagger} (0,0)e^{i\omega_{\bf k} t_1 }\nonumber \\
\nu(t_1 )=i\sum_{\bf k} g({\bf k}) a_{\bf k} (0,0)e^{-i\omega_{\bf k} t_1}
\label{eq41b}
\end{eqnarray}
Note that when calculating the quantum mean value with an initial thermal state, the terms proportional to $\nu$ and $\nu^\dagger$ vanish, so that the exact evolution equation of $\langle A(t)\rangle$ only depends on the correlation function (\ref{chapuno325}). A similar calculation for two time correlation functions of system observables $A$ and $B$ leads to the form
\begin{eqnarray}
&&\frac{dA(t_1 )B(t_2 )}{dt_1}=i [H_S (t_1 ),A(t_1 )]B(t_2 )
\cr&-&\nu^{\dagger}(t_1 )[L(t_1 ),A(t_1 )]B(t_2 )+[L^{\dagger}(t_1 ),A(t_1 )]B(t_2 )\nu(t_1 )\nonumber\\
&-&\int_0^{t_1 }d\tau \alpha^* (t_1 -\tau)L^{\dagger}(\tau)[L(t_1 ),A(t_1 )]B(t_2 )\nonumber\\
&+&\int_{t_2}^{t_1 }d\tau \alpha(t_1 -\tau)[L^{\dagger}(t_1 ),A(t_1 )]L(\tau)B(t_2 )\cr
&+&\int_0^{t_2} d\tau \alpha(t_1 -\tau)[L^{\dagger}(t_1 ),A(t_1 )]B(t_2 )L(\tau).
\label{eq45}
\end{eqnarray}
The evolution of the quantum mean value $\langle A(t_1 )B(t_2 )\rangle$ is again obtained by computing the trace with the total initial state on both sides of the former expression, finding out that the resulting exact equation only depends on the correlation function. A generalization to an N-time correlation function can be found in \cite{alonso2007}.
\section*{Appendix B: Independent environment limit}
To illustrate the idea presented in Section (\ref{indep}), let us assume the Caldeira and Legget model for the spectral density, which gives a good approximation of the spectral densities of different types of environment in the short frequency limit \cite{caldeira1983,wallsbook,leggett1987},
\begin{eqnarray}
J(\omega)=\alpha\omega_c^{1-s}\omega^s\theta(\omega_c-\omega).
\label{chapuno41}
\end{eqnarray}
Here, $0<s<1$ in the \textit{sub-ohmic} case, and $s>1$ in the \textit{super-ohmic}
where $\omega_c$ is a frequency cut. To reproduce such spectral density, we assume a dispersion relation of the form $\omega(k)=Ak^p$, where $A$ and $p$ are constants to be chosen as convenient. In terms of this, the corresponding density of states is $\rho_\textmd{DOS}(\omega)=\large|\frac{\partial k}{\partial \omega}\large|=\frac{A^{\frac{1-p}{p}}}{p}\omega^\frac{1-p}{p}$. Considering $g(k)=g$, i.e. a homogeneous coupling Hamiltonian as in (\ref{h1EM}), we find that
\begin{eqnarray}
J(\omega)=g^2\rho_\textmd{DOS}(\omega)=g^2\frac{A^{\frac{1-p}{p}}}{p}\omega^\frac{1-p}{p}.
\end{eqnarray}
Hence, to reproduce an ohmic spectral density ($s=1$), we need to chose $p=1/2$, so that $(1-p)/p=1$, and the constant $A=1/2$. Similarly, a sub-ohmic spectral density of the form ($s=1/2$) will require choosing $p=2/3$ and $A=4\omega_c/9$, and a super-ohmic like $J(\omega)=\frac{\omega^{2}}{\omega_c}$ will require $p=1/3$ and $A=\sqrt{\frac{1}{3\omega_c}}$. In all cases, the coupling should be chosen as $g=\sqrt{\alpha}$.
In general, for a spectral density with $s$, we need $p=1/(s+1)$ and $A=(\frac{\omega_c^{1-s}}{s+1})^{1/s}$.
|
\section*{Abstract}
We compile spectroscopic abundance data from 84 literature sources for
50 elements across 3058 stars in the solar neighborhood, within 150 pc
of the Sun, to produce the {\it Hypatia Catalog}. We evaluate the
variability of the spread in abundance measurements reported for the same star by different surveys.
We also explore the likely association of the star within the Galactic disk,
the corresponding observation and abundance determination methods
for all catalogs in Hypatia, the influence of specific catalogs on the overall abundance trends,
and the effect of normalizing all abundances to the same solar scale.
The resulting large number of stellar abundance determinations in the
{\it Hypatia Catalog} are analyzed only for thin-disk stars with
observations that are consistent between literature sources.
As a result of our large dataset, we find that the stars in the solar neighborhood may be reveal an asymmetric abundance distribution, such that a [Fe/H]-rich group near to the mid-plane is deficient in Mg, Si, S, Ca, Sc II, Cr II, and Ni as compared to stars further from the plane.
The {\it Hypatia Catalog} has a wide
number of applications, including exoplanet hosts, thick and thin disk stars,
or stars with different kinematic properties.
\section{Introduction}
\label{s.intro}
One of the primary tools to understanding the history of the solar neighborhood,
and more generally the Milky Way, is the chemical composition of stars.
From the initial efforts of \citet{Russell:1929p2477}, \citet{suess_1956_aa},
and \citet {Bidelman:1960p2486}, to the more recent works of \citet{Anders:1989p3165},
\citet{Edvardsson:1993p2124}, \citet{Bensby:2005p526}, \citet{Valenti:2005p1491},
\citet{Asplund:2009p3251}, and \citet{Lodders:2009p3091}, compilations of
stellar abundances provide an overall picture of the chemical evolution
of the solar neighborhood. Notable results obtained over the past few decades include
correlations of metallicity with age and Galactocentric distance, and whether the
Sun is suitably ``average''
\citep{eggen_1962_aa,twarog_1980_aa,Feltzing:2001p867,robles_2008_aa,robles_2008_ab}.
Trends in the elemental abundances, and a limited number of isotopic abundances,
relative to iron have also been observed over the whole metallicity range of the Galactic disk
\citep{Venn:2004p1483,Soubiran:2005p1496}. For example, oxygen and
the other $\alpha$-chain elements, relative to iron, vary systematically from
being overabundant at [Fe/H] $\ltaprx$ 1.0 dex to roughly solar at
[Fe/H] $\approx$ 0.0. This decrease is widely taken to be caused by the
contributions of supernovae Type Ia (SN Ia) in
an average, well-mixed interstellar medium
\citep{truran_1971_aa,tinsley_1980_aa,matteucci_1986_aa,lambert_1989_aa,
Wheeler:1989p3310,Timmes:1995p3197, goswami_2000_aa,Gibson:2003p2583,
kobayashi_2006_aa, Krumholz:2007p3338,Prantzos:2008p2591,romano_2010_aa,
kobayashi_2011_aa}.
Another tool to interpret the history of the solar neighborhood (taken
throughout this paper to be stars within 150 pc of the Sun) is the
theory of stellar evolution, nucleosynthesis, and chemical evolution.
By quantifying the ejecta from stars, this history can be
reconstructed using theoretical models
\citep{burbidge_1957_aa,cameron_1957_aa,Woosley:1995p3481,
Thielemann:1996p3396,meynet_2002_aa,Siess:2002p3399,Ventura:2002p3389,
Limongi:2003p3406,karakas_2007_aa,jose_2011_aa}.
These models account for the
initial mass function, star formation rate, stellar yields, inherited
composition from the local interstellar medium (ISM), and are one of
the keys to quantifying the formation of local solar neighborhood
stars. Their results help provide important constraints on chemical
evolution of the solar neighborhood, the Galactic disk, and other
galaxies.
As the number of spectroscopic surveys of stars in the solar
neighborhood increases, it has become tradition for authors of
abundance surveys or chemical evolution models to
compare their relative abundances to benchmark data sets
for verification or validation. Typically, this involves comparing
to \citet{Edvardsson:1993p2124}, \citet{Reddy:2003p1354},
\citet{Bensby:2005p526}, or \citet{Valenti:2005p1491}. However, the
manner by which these comparisons are conducted varies
drastically. Some authors provide statistical evaluations such as mean
differences and standard deviations, some compare a few ``typical''
stars, and others graphically juxtapose entire catalogs. While there
are certainly correlations between published data sets, there has been
little discussion of the nuances, random uncertainties, and systematic
biases of the compared data sets. It is, however, these idiosyncrasies
that make interpreting trends between abundance catalogs challenging.
It is difficult and rare for one survey to systematically observe a
large number of nearby stars and provide abundance determinations for
a wide variety of elements. For example, \citet{Valenti:2005p1491}
reported the relative abundance in 1040 stars for only five elements, including
iron. Alternatively, \citet{Reddy:2003p1354} analyzed
spectra for 27 elements, but their study had only 181 stars. To
achieve the most complete coverage of the solar neighborhood, the
relative abundances from known literature sources must be combined.
Such compilations have been undertaken by, for example,
\citet{Venn:2004p1483} and \citet{Soubiran:2005p1496}, with the
amalgamation of thirteen and eleven published catalogs, respectively. \citet{Venn:2004p1483} contrasted the elemental abundance trends found in the nearby dwarf spheroidal galaxies with the overall trends found in the Milky Way in order to investigate whether, for example, the Milky Way halo could have been built from such systems.
In a vein similar to our own, \citet{Soubiran:2005p1496} analyzed the abundance trends between $-1.3 <$ [Fe/H] $< +0.50$ for thin and thick disk stars. In both cases, the authors compiled detailed element abundances for a large number of stars in order to better understand overarching trends within the Milky Way.
The primary purpose of this paper is to present our {\it Hypatia Catalog}
-- so named after one of the first female astronomers who lived around 400 A.D..
The {\it Hypatia Catalog} (hereafter, Hypatia) is an unbiased compilation of
spectroscopic abundance determinations from 84 literature sources for
50 elements across 3058 stars within 150 pc of the Sun. For the purposes
of this paper, we show that the stellar abundances within Hypatia reflect the
zeroth order trends consistent with published literature and mean galactic chemical
evolution models. In \S \ref{s.comp}, we
discuss the collation of the data, the inherent challenges
in combining different data sources.
In \S \ref{s.analysis} we discuss the analysis of the stars in the Hypatia, whether they are thin- or thick-disk stars, and our attempts to mitigate some of the challenges in compiling such a large dataset. These adjustments to the data were only
for the purposes of analysis and are not reflected in the published catalog.
In \S \ref{s.struct}, we describe more thoroughly how Hypatia was compiled.
In \S \ref{s.cat}, we describe some of the details of the {\it Hypatia Catalog} as well
as the abundance trends seen in Fe.
In \S \ref{s.alpha}, we present the abundance trends found in the $\alpha$-elements $-$
C, O, Mg, Si, S, Ca \& Ti $-$ in the solar neighborhood.
In \S \ref{s.odd}, we describe the abundance trends for the odd-Z elements, namely
Li, N, Na, Al, P, K \& Sc.
The trends found in the iron-peak elements (V, Cr, Mn, Co \& Ni) are discussed
in \S \ref{s.ironpeak} while those elements that are beyond the iron-peak (Cu, Zn, Sr, Y, Zr, \& Mo)
are given in \S \ref{s.beyond}. Finally, the neutron-capture elements (Ru, Ba, La, Ce, Pr, Nd, Sm,
Eu, Gd, Dy, \& Pb) and their abundances are presented in \S \ref{s.neutron}.
Finally, in \S \ref{gaps}, we analyze an anomaly in the abundance measurements of a number of different elements that presented ``gaps" in [X/Fe] with for all [Fe/H] and discuss the physical implications.
\section{Compilation of the Hypatia Catalog}
\label{s.comp}
In this section we
detail the scope of Hypatia and the manner in which datasets were chosen to be included.
We also go into a more critical analysis of the individual works from which the abundance data were determined, given in a tabular format in Table \ref{tab.long}.
\subsection{Abundance Datasets}
Numerous studies have analyzed the photospheres of stars in the solar
neighborhood using photometric and spectroscopic techniques.
Photometric investigations have treated a much larger number of stars
relative to spectroscopic methods. However, photometric studies
generally yield one global metallicity parameter, [Fe/H] $= \log \,(
N_{Fe} / N_H)_{*} - \, \log \,( N_{Fe} / N_H )_{\astrosun}$, with
units in dex, where $N_{Fe}$ and $N_H$ are the number of iron and
hydrogen atoms per unit volume, respectively. Despite the smaller
number of stars analyzed with spectroscopy, the additional element
abundances allow assessment of not just the overall metallicity, but
the full chemical compositional range and evolution. Therefore, we
have chosen to focus on published spectroscopic abundance catalogs.
We compiled Hypatia with the spectroscopic abundance determinations of
50 element abundances for 3058 unique stars from published catalogs.
Our exhaustive literature search considered all abundance
determinations, of which we are aware, for main sequence
F/G/K/M-type stars within 150 pc of the Sun. Table \ref{tab.hyp} shows a sample of the {\it Hypatia Catalog}
which includes stellar HIP/HD/BD names, spectral type, distance,
position, and the compiled abundances as given by each catalog, with
reference. The complete catalog is given in the electronic version of
this paper and the reduced catalog (per the discussion in \S \ref{s.analysis} and \S{s.cat}) will be made available on Vizier. Efforts were made to include literature sources with
abundance measurements for local stars published before the original submission of this (and any following) paper; any exclusion was not intentional. In addition, Hypatia has and will continue to be updated as newer and more precise abundances are determined.
While these updates will be announced in subsequent papers, Hypatia will also be made available
as an independent online database such that the community will have access to the most up-to-date data.
Therefore, if a star within the solar neighborhood was
measured for abundances other than iron, and it is 150 pc of the Sun, it will be incorporated into
Hypatia.
The data sets that are contained in Hypatia are listed in
Table \ref{tab.cat}, along with the number of stars meeting the above
criteria and the element abundances determined therein.
Throughout the paper, we give a more detailed description of each literature
source and their method for determining stellar abundances.
A histogram of the number of stars measured for each element in Hypatia is
shown in Fig.\,\ref{fig.hist}. All 3058 stars have a spectroscopically determined [Fe/H].
The next most frequently measured elements in Hypatia are Si (2320 stars),
Ti (2237 stars), Ni (2225 stars), and O (2125 stars). There are only 33
stars in the solar neighborhood for which [Ru/H] has been measured and
only 20 stars for [P/H]. Fig.\,\ref{fig.hist} also shows the relative
paucity of stars in the solar neighborhood that have had their
nitrogen, magnesium, and sulfur abundances determined,
showing a clear direction for future studies of under-explored elements.
This is primarily due to having too few absorption lines, or lines
that are too weak to separate from the continuum in the optical
spectrum. The {\it Hypatia Catalog} also has abundance measurements for 13
singly ionized elemental species. While some catalogs measured only one ionization
state when reporting an abundance determination, a number of catalogs
combined the abundances from multiple ionization states. In Hypatia,
an abundance of [X/Fe] means that a catalog measured the neutral
state, a combination of neutral and ionized state(s), or it was not
specified. Whenever a catalog specifically mentioned it was only
measuring the singly ionized state, we write [X II/Fe].
\begin{figure}[ht]
\centerline{\includegraphics[height=2.5in]{f1.pdf}}
\caption{
Number of stars in the {\it Hypatia Catalog} with measured abundances
for 46 different element species.
}\label{fig.hist}
\end{figure}
\subsection{Comparison of Datasets}
It was a large endeavor to compile and merge the 84 datasets in Hypatia, using services such as VIZIER, ADS, and SIMBAD as well as often times going to the original source in order to transcribe tables not available online. However, we recognize that it is naive to collect data without critically analyzing the individual works. Table \ref{tab.long} presents a comprehensive inventory of the datasets with 20 or more stars found in Hypatia, including the telescopes and spectrographs, resolution, signal-to-noise (S/N), and wavelength range. We also documented the methods used by the individual authors to determine the stellar abundances, such as the stellar atmosphere models, equivalent width measurement techniques, whether they
used iterative force-fits of observed and synthesized equivalent widths (i.e. curve-of-growth) or
spectral fitting procedures, the solar abundances employed, and the number of Fe I and Fe II lines. Data were compiled with as much information as provided by the authors in their papers or subsequent references. Care was also taken to cite the appropriate sources, when available. Any error or exclusion was not intentional.
Looking at the catalogs given in Table \ref{tab.long}, it is clear that a wide variety of telescopes and spectrographs have been employed in order to determine stellar abundances in the local neighborhood. The telescope mirror sizes have a range from 1 m at the Cerro Tololo Inter-American Telescope \citep[CTIO,][]{Laird:1985p1923} to the 10 m telescope at Keck using the High Resolution Echelle Spectrometer \citep[HIRES,][]{Boesgaard11}, where the average aperture size is 3.4 m for all of the telescopes listed in Table \ref{tab.long}. A total of 18 of the 58 datasets combine spectra from two or more telescopes when determining their abundances. The resolution from the instruments varies from $\Delta \lambda / \lambda$ $\approx$ 30\,000 \citep{Gebran:2010p6243, Kang11} to 120\,000 \citep{Ramirez:2007p1819,Ramirez:2009p1792}, where \citet{Gilli:2006p2191,Ramirez:2007p1819,Ramirez:2009p1792} cite a range in $\Delta \lambda / \lambda$ $\approx$ 60\,000 in their spectra alone, which was combined from multiple telescopes/spectrographs. Signal-to-noise was in general 100--200 per spectral pixel, where the largest span (90--1350 per pixel) was encompassed by the spectra amalgamation in \citet{Ecuvillon:2004p2198,Ecuvillon:2006p8109}.
The methods employed in order to determine the abundances from the spectra were also rather diverse. The preferred programs for determining stellar atmospheres were MARCS per \citet{Gustafsson:1975p4658} and ATLAS by \citet{Kurucz1993}. However, a number of alternative routines were also employed, such as EAGLNT \citep{Edvardsson:1993p2124} and MAFAGS \citep{Fuhrmann:1997p6740}. Of the 58 datasets in Table \ref{tab.long}, a total of 48 of them determined equivalent widths of the absorption lines, for example using IRAF splot, ARES, or WIDTH9 \citep{Kurucz1993}, and then used the curve-of-growth method, per MOOG \citep{Sneden:1973p6104} or Uppsala EQWIDTH, to calculate the
majority of
element abundances.
Conversely, there were 10 datasets who used spectral fitting as their predominant method of determining stellar abundances.
Many of the datasets in the {\it Hypatia Catalog} used \citet{Anders:1989p3165} and \citet{Grevesse:1998p3102} for their solar abundance scale, however, many also calculated their own. As discussed in \S \ref{s.spread}, we have divided out the solar abundance used in each individual dataset and instead applied the abundance scale by \citet{Lodders:2009p3091}.
Finally, while most of the catalogs that are incorporated into Hypatia published their line lists, not all of them measured the same element abundances, making a standardized comparison difficult. However, as mentioned in \S \ref{s.comp}, it was a requirement that any dataset added into Hypatia have [Fe/H] abundance measurements. Therefore, we have included the number of Fe I and Fe II lines in the last column of Table \ref{tab.long}. Through work as a part of the Stellar Stoichiometry: Workshop Without Walls held at Arizona State University during April 2013 (Hinkel et al. in prep),
it was confirmed that the number of iron lines measured, below a certain threshold, drastically affects the determined [Fe/H] abundance regardless of the abundance measurement technique employed. This is of particular interest given that the number of Fe I / Fe II lines varies from 20 / 1 \citep{Jonsell:2005p1298} to 450 / 25 \citep{Luck:2005p1439} -- ignoring those catalogs for which a definite count could not be made, namely \citet{Gustafsson:1999p8407} and \citet{Thevenin1999}.
From Table \ref{tab.long} we have found that a number of datasets included in the {\it Hypatia Catalog} use different telescopes/spectrographs at varying resolutions with a range in S/N, along with an assortment of prescriptions for modeling the stellar atmospheres and methods for measuring a variety of absorption lines in the spectra to determine abundances. However, we opted to avoid any sort of hierarchy on which to place the datasets, for fear that our ``choices" may be inadvertently arbitrary or without the scientific rigor demonstrated in \citet{Lebzelter12}. We also found that scaling the abundances based on their stellar parameters was not viable per \citet{Ramirez12}. Instead, we examine the uniformity of the elemental abundances on a star-by-star basis between the datasets in \S \ref{s.spread} and exclude inconsistent measurements based on standard error in our further analysis.
\section{Analysis of Hypatia Stars}
\label{s.analysis}
Here we examine more closely the stars within the {\it Hypatia Catalog}, for example whether they are likely to be from the thin or thick disk population of the Galaxy. We also study the discrepancy in abundances when multiple datasets measure the same element within the same star.
As a result of the following analysis, we are able to define a subset of abundances within the {\it Hypatia Catalog} that is both homogenous and
robust.
\subsection{Thin vs. Thick Disk Stars}
\label{s.thinthick}
We separated the stars in the {\it Hypatia Catalog} into the Galactic components, for example thin-disk, thick-disk, or halo stars, based on their kinematics. Since chemical differences between the stellar populations are still debated, and chemical trends are the topic of our investigation, we decided that a conservative kinematic approach was the most appropriate. We used the updated and extended $UVW$ space velocities as determined for the {\it Hipparcos Catalog} by \citet{Anderson12}. The local standard of rest used in their calculations was $(U_0, V_0, W_0) = (-14.0, -14.5, -6.9)$ km\,s$^{-1}$ \citep{Francis09}. We followed the prescription in \citet{Bensby:2003p513} in order to determine the probability that the Hypatia stars belonged to one of the Galactic components. By assuming the space velocities of the three stellar populations, as well as the number densities of the stellar components in the solar neighborhood, we were able to get a probability that a star belonged to a certain population. We assumed that 18\% of the stars in the solar neighborhood originated in the thick disk, per \citet{Adi13}. This local relative stellar density does not agree with \citet[][and references therein]{Bensby:2003p513} who assumed a 6\% thick-disk population. However, it is lower than the 25\% adopted in \citet{Mishenina:2004p1360} who later estimated an expected range of 2-15\% with a larger sample size, similar to that found in \citet{Kordopatis11}, and larger than that noted in \citet{Reyle01}. Therefore, we found 18\% to be a good median that was more conservative and agreed with the more recent studies. We did adopt the Gaussian distributions used in \citet{Bensby:2003p513}.
According to our probabilities, 2537 stars in Hypatia are thin-disk stars, 369 are thick-disk stars, 0 were from the halo, and 163 did not have space velocities according to \citet{Anderson12}. Therefore, we find that 12\% of the {\it Hypatia Catalog} stars are likely from the thick-disk, as illustrated in red in the Toomre diagram
\citep[A. Toomre 1980 private communication with][]{Sandage1987}
in Fig. \ref{uvw} (left). Our local stellar density calculations are consistent with the literature and lie within the expected range estimated by \citet{Mishenina:2004p1360}. {\bf We did not include any of the
probable
thick-disk stars in our chemical abundance analysis in \S \ref{s.cat} -- \S \ref{s.neutron}.}
A number of people have noted, for example \citet{Adi13, Bensby:2003p513}, and references therein, that the Galactic components of the solar neighborhood may be discerned via a combination of both kinematics and chemical abundances. In their paper, \citet{Adi13} described a population distinction occurring in an [$\alpha$/Fe] versus [Fe/H] diagram, where ``$\alpha$" refers to the average of the Mg, Si, and Ti abundances. More pointedly, a region devoid of abundances measurements should occur approximately at [$\alpha$/Fe] $\approx$ 0.2-0.3 dex, with a possible ``knee" at [Fe/H] $\approx$ -0.3 dex (see their Fig. 1). We have recreated the \citet{Adi13} figure in Fig. \ref{uvw} (right) where the thin-disk stars are light yellow circles and the thick-disk stars are red triangles. The two populations in the {\it Hypatia Catalog} appear relatively well-mixed with respect to [$\alpha$/Fe],
with few thick-disk stars that have lower [$\alpha$/Fe] and higher [Fe/H].
Our findings mimic the trends seen in \citet{Reddy:2006p1770, Pritzl05, Tolstoy03, Edvardsson:1993p2124}, who also analyze the trends seen in [$\alpha$/Fe] compared to [Fe/H] for thin- and thick-disk stars. They also do not show the same ``knee" or empty region displayed in \citet{Adi13}.
\begin{figure}[ht]
\centerline{\includegraphics[height=3.0in]{f2.pdf}}
\caption{A Toomre diagram showing all of the stars in the {\it Hypatia Catalog} (left), where the stars most likely to be from the thick-disk are in red and those from the thin-disk are in black. To the right, a similarly color-coded plot recreating Figure 1 seen in \citet{Adi13}, where they found the two populations separated by a gap in [$\alpha$/Fe], which was not seen using the Hypatia stars.}
\label{uvw}
\end{figure}
\subsection{Spread in the Elements}
\label{s.spread}
Collecting abundance determinations from multiple authors over about a
25 year time span means at least the following differences between
data sets:
instrument zero points,
resolution of the spectra,
signal-to-noise ratios,
oscillator strengths,
line lists,
equivalent widths,
number of ionization stages used,
LTE or non-LTE analysis,
converged solar atmosphere models,
curve-of-growth or spectral fitting,
curve-of-growth program used, and
adopted solar abundances.
All of these factors may introduce systematic and stochastic
differences between data sets. For example, Fig. \ref{spread} (top) shows
the abundance measurements for six elements within five Hypatia stars.
The circles are as labeled with the element name while all triangles designate [Fe/H], each with
respective error bars from the catalog from which it was measured. The
variation between catalogs per element, the largest of which we call
the {\it spread},
has a mean of 0.14 dex and a median of 0.11 dex for all elements in all stars in Hypatia. These values are on par if not larger than the error bars for most elements.
In addition to the abundances, stellar parameters
may also suffer from inherent issues, depending on the reduction procedure. \citet{Torres12} analyzed three methods
for determining T$_{eff}$, $\log$(g), and [Fe/H] in order to find the systematic differences between
the techniques. They utilized both ``constrained" spectroscopic determinations of the stellar parameters, by making use of the normalized semimajor axis $a/R_{*}$, as well as ``unconstrained" determinations. Through these comparisons, they found that in the case of Stellar Parameter Classification (SPC) and Spectroscopy Made Easy (SME), both of which employ spectral synthesis and global $\chi^2$ minimization, the surface gravity was strongly correlated to the T$_{eff}$ and [Fe/H] measurements (see their Figures 2 and 4). In other words, when the $\log$(g) values are larger, the T$_{eff}$ and [Fe/H] measurements increase due to the degeneracy between the stellar parameters. Additionally, they noted that the variation for each respective parameter between the ``constrained" and ``unconstrained" determinations increased as effective temperature increased (their Figure 6).
Within the {\it Hypatia Catalog}, Fig. \ref{spread} shows the variation of T$_{eff}$ and $\log$(g) (middle and bottom,
respectively) among many of the included datasets, determined for the abundances given at the top of the figure (when reported).
For T$_{eff}$, there is a
spread of $\sim$200 K, while $\log$(g) can differs by a factor of $\sim$ 0.3.
The overall abundance determinations are sensitive to the adopted
stellar values, especially $\log$(g).
\begin{figure}
\centerline{\includegraphics[height=6in]{f3.pdf}}
\caption{
The representative \textit{ spread} (left) from different catalogs in the
un-normalized abundance determinations (see text) for 6 element ratios
for 5 stars, with quoted catalog errors such that the quoted
uncertainty of individual measurements is less than the spread. The element ratios
[X/H] for a given star are denoted by circles and the corresponding
[Fe/H] abundances are shown as triangles.
In the middle, the T$_{eff}$ values for the same stars and $\log$(g) are
at the bottom.
}
\label{spread}
\end{figure}
There is an accumulation of systematic and stochastic differences in
the abundance measurements compiled to form Hypatia.
Other authors have noted the difficulties in comparing different
catalogs \citep[e.g.][]{Feltzing:1998p886,Bond:2006p2098}, but only recently
have they tried to overcome the challenges.
\citet{Ramirez12} analyzed the lithium abundances from nearby stars, supplemented by abundances from 7 other catalogs. They attempted to scale the abundances within the catalogs by taking into account the differing stellar parameters (namely, T$_{eff}$ and $\log (g)$). However, the abundances did not vary with the stellar parameters in a consistent manner, not linearly or even as a standard function. Because stellar atmospheric models determine a wide variety of parameters, we agree with \citet{Ramirez12} that attempting to correct for differing stellar parameters between catalogs will not be scientifically valuable. Therefore, we have decided to use a method similar to \citet{Roederer13}. He combined 54 nearby stellar abundance catalogs and only used the highest quality abundances for repeat observations, making no attempt to correct for differing stellar parameters. His data was therefore unbiased towards one specific dataset, unlike \citet{Ramirez12} who opted to normalize the other datasets to their own.
Due to the range of possible issues, we have not included T$_{eff}$ and $\log$(g) for the stars in Hypatia.
We attempted to make the various catalogs in Hypatia more copacetic by putting all the different
measurements on the same solar abundance scale,
a correction that was also employed in \citet{Hinkel13},
see Table \ref{tab.long}, Column 9.
While some literature sources adjust their atomic data in order for their results to match certain
solar values, this correction takes place before the stellar-to-solar comparison and therefore before the
solar normalization, or re-normalization.
Therefore, normalization to the same solar scale is the only correction
available to us that helps make the data more comparable and does not involve recalculating the abundance determinations
from every dataset.
For example, per Table \ref{tab.hyp}, \citet{Valenti:2005p1491}
reported HIP 400 to have [Ti/H] = -0.28 dex. In their paper, they
cite \citet{Anders:1989p3165} as the source of their solar abundances,
where $\log \epsilon(Ti)$ = 4.99. To calculate the re-normalized abundances
according to the solar measurements by \citet{Lodders:2009p3091}, where $\log \epsilon(Ti)$ = 4.93, then
[Ti/H] $= -0.28 + 4.99 - 4.93 = -0.22$ dex.
For all of the abundances within each star in the {\it Hypatia Catalog}, we found that the average difference
between the abundance determination before and after the re-normalization was 0.06 dex, with a median
of 0.04 dex. These variations in abundance are larger than the quoted error for many of the elements determined by the literature datasets. On the other hand, the average and median spread found in Hypatia only changed by 0.01 dex as a result of the re-normalization. In other words, the choice of solar abundance scale significantly affects the element abundance measurement, however, re-normalizing all of the datasets to the same solar abundance scale does not reduce the spread in the data.
To be as consistent as possible, for all Hypatia calculations hereafter, we retain the element
abundance values using the \citet{Lodders:2009p3091} solar abundance
renormalization.
\section{The Structure of the Hypatia Catalog}
\label{s.struct}
The {\it Hypatia Catalog} was compiled using PYTHON 2.7.3 with the following packages: MATPLOTLIB 1.3.1, NUMPY 1.8.0, and ATPY 0.9.6. Abundances tables (that measured both [Fe/H] and at least one other element [X/Fe]) were either downloaded from VIZIER, converted to machine-readable formats from the \LaTeX manuscripts supported by ADS, or were manually transcribed. The stellar names used in each individual table were then matched to the Hipparcos naming-scheme, which was chosen in order to provide continuity, incorporate as many stars as possible in the solar neighborhood, and provide a variety of stellar parameters (see Table \ref{tab.hyp}). Datasets were only included in Hypatia if at least one star was within 150 pc of the Sun with a F, G, K, or M spectral type. In addition, we did not include abundances in our analysis that were determined using non-local thermodynamic equilibrium (NLTE) approximations.
If a dataset had stars that met the above criteria, than the element names, abundance values, and the literature source are recorded for each star, which is then incorporated into Hypatia. However, if a star within the dataset was found to likely originate from the thick disk or halo (per our discussion in \S \ref{s.thinthick}), that star was not incorporated in the compilation. Meaning, while a dataset may have been included, all of their stars are not necessarily part of the Hypatia Catalog. During this process, we also renormalized all of the stellar abundances, using \citet{Lodders:2009p3091} as the standard. Once the renormalized abundances from all 84 datasets are compiled into Hypatia, then we are able to analyze the data. An element is first chosen to study, for example, sodium. Every thin disk star with a sodium abundance measurement is determined. In the cases where multiple datasets measured the same element abundance in the same star, so as not to favor any catalog like \citet{Roederer13}, the median value for those measurements is used. However, we found that if the discrepancy between catalog measurements is too large, the median abundance value was unreliable. Rather than preferentially choose one catalog over another, we opted to eliminate those stars that do not have consistently measured element abundances from our analysis, in lieu of those that are more uniform and of higher quality. Therefore, {\bf any star with a spread in either [X/H] or [Fe/H] larger than the respective error bar was not included in the following (plotting) analysis. This cutoff value was used because it highlights those elements that are not well agreed upon by multiple literature sources.}
The total number of well-understood element abundances that remain for thin disk stars (see \S \ref{s.thinthick}) with spread per element less than error bar are listed in Table \ref{tab.skip}. Reduced abundances for these stars will be made available on Vizier.
In order to determine the representative error associated with each of the elements, we recorded the error as reported by each of the datasets from which that element was measured and averaged them together. Very few surveys determined star-by-star error values for the abundances, forcing us to also use a more general error bar per elements. In those cases where there were abundance errors per star, we took the average of all of the individual errors. While this method may mask some of the more precise abundance measurements with smaller associated errors, we found that it was better to take the more conservative approach by over-estimating the error, rather than under-estimating.
Going back to the example, for the case of sodium, there are 907 thin disk stars with one sodium abundance each that either came from 1) a single dataset or 2) the median value of multiple datasets measurements that have small variation between them. Because every star and every element abundance measurement are not created equal, this means that we are not plotting the same stars for every element. Instead, each element within each star is analyzed on a case-by-case basis in order to ensure that the quality of the data is high for our analysis, such that there were no compilation errors or idiosyncrasies. To create the [X/Fe] vs. [Fe/H] plots prevalent in \S \ref{s.alpha}-\ref{s.neutron}, we binned each of the stars according to distance from the Sun, as calculated from the RA, Dec, and parallax angles associated with the updated Hipparcos catalog \citep{Anderson12}.
Quantifying systematic errors that could result from
instrumental, atomic database, or stellar atmosphere models is beyond
the scope of this paper, but we encourage the community to undertake
such verification and validation studies. Therefore, we have decided to publish the
{\it Hypatia Catalog}, see Table \ref{tab.hyp}, containing only the original data by the
literature authors, without any of our alteration or combination. In this way, others
may use the compiled data to perform a more selective analysis as they see fit.
\section{Abundances in the Hypatia Catalog}
\label{s.cat}
Combining 84 data sets that span about 25 years means there will be a
{\it spread} among reported values for any element in many stars. We
have have taken a many steps to address the issues to help make results
generated with Hypatia meaningful and physical. We (a)
exclude any probable thick-disk stars per the \citet{Bensby:2003p513} method, illustrated
in the Toomre diagram (Fig. \ref{uvw}, left) in \S \ref{s.thinthick}
(b) attempt to minimize the spread by renormalizing the abundances to
a standard solar abundance scale in \S \ref{s.spread}; (c) exclude stars
with a large spread greater than the error bar per \S \ref{s.struct}; and (d) choose to use the median value of the spread
avoid specific catalog bias when representing the abundance of a star in \S \ref{s.struct}.
{\bf
In this way, we only analyze
thin-disk stellar abundances that are consistently measured between literature sources.
}
The elemental abundances were then plotted in the traditional [X/Fe] versus [Fe/H] plane in Figs. \ref{ferad}-\ref{ndeu}.
Representative error bars, compiled from the quoted observational uncertainties given by each literature source, for each element are placed in the upper right corners of each figure. We also show the overall trends in
[X/Fe] vs. [Fe/H] by plotting 1$\sigma$ quantile regression.
We also analyze the data for trends in radial distance, z-height above the galactic disk, and
directionality per the galactic center and anti-center. We have included a handful of these
plots for a variety of elements to demonstrate any physical trends that may
be present. There are also a small number
of paired plots with abundances before and after the
\citet{Lodders:2009p3091} renormalization. Literature sources that
have contributed to the {\it Hypatia Catalog} are discussed
throughout.
\subsection{Iron}
\label{s.iron}
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f4.pdf}} \caption{
Median [Fe/H] ratio for
thin-disk, consistently measured
stars in Hypatia as a function of radial distance
from the Sun. Horizontal error bars along the top corresponding to
the error in parallax angle used to calculate the distances.
}
\label{ferad}
\end{figure}
The principle energy sources for most stars are hydrogen burning via the pp-chain or the CNO-cycle,
or $\alpha$-chain burning. These energy sources underlie the processes
through which most of the naturally occurring elements are
created \citep{burbidge_1957_aa,Woosley:1995p3481,Thielemann:2002p3625,jose_2011_aa}.
Iron is relatively easy to measure within stars due to the large number of absorption lines in the optical regime. Both SN Ia and core-collapse supernovae (SN II, SN Ib/c)
produce iron, on different timescales as well as in different and debated amounts
\citep{chiappini_1997_aa, Thielemann:2007p4473, Prantzos:2008p2591}.
And, due to the multiple production sites, the mean trend of iron increases monotonically
in time within the ISM and thus it can act as a chronological indicator of nucleosynthesis
\citep{Wheeler:1989p3310,pagel_1997_aa,vangioni_2011_aa}.
Therefore, as [Fe/H]
increases, so does the general timeline of chemical evolution -- where
we expect to see contributions from core-collapse supernovae
for low values of [Fe/H] and
the effects from SN Ia at higher values
\citep{Wheeler:1989p3310,matteucci_2001_aa,Gibson:2003p2583,chiappini_2011_aa}.
However, stars can migrate or scatter into or out of the solar
neighborhood, and different galactic populations can have different
star formation histories. In this case, [Fe/H] does not necessarily
represent the same timeline, which may introduce some ambiguity in
using [Fe/H] as a chronometer
\citep{wielen_1996_aa,gratton_1996_aa,sellwood_2002_aa,haywood_2008_ab,prantzos_2011_ab}.
Fig. \ref{ferad} shows the median values of [Fe/H] reported for
1713 stars in Hypatia, excluding those with a spread larger than the error
(see \S \ref{s.spread}),
with respect to the radial distance from the
Sun. The horizontal error bars along the top are 0.32 pc at 20 pc, 5.1 pc
at 80 pc, and 16.3 pc at 140 pc, showing how the fractional uncertainty
in parallax angle affects the uncertainty in the distance calculation.
Within our solar neighborhood's radius of 150 pc,
there is a
relatively constant scatter in [Fe/H] at any distance, which may be due to a
similar stellar origin or homogeneous mixture. The scatter in [Fe/H] spans $\approx$
1.5 dex, although the vast majority of the stars lie within [-0.2, 0.5]
and are mostly within 80 pc of the Sun. There are a few thin-disk stars, however, that
are a relatively low in [Fe/H], with abundances $<$ -1.0 dex.
Given that 150 pc
is small on galactic scales, near-solar values are to be expected.
\section{Chemical Abundances of $\alpha$-Elements: \\ C, O, Mg, Si, S, Ca, \& Ti}
\label{s.alpha}
CNO nuclei are among the most abundant elements in the solar
neighborhood
\citep{Anders:1989p3165,Lodders:2009p3091,Asplund:2009p3251}. They
are important in stellar interiors as opacity sources
\citep{Iglesias:1996p3447}, as energy producers through the CNO cycle
\citep{Bethe:1938p3459}, and are essential building blocks of
terrestrial biochemistry \citep{Pace:2001p3477}.
For consistent comparison, we have plotted all of the $\alpha$-elements using the same x- and y-axis scales, such that [Fe/H] = [-0.65, 0.65] and [X/Fe] = [-0.7, 0.7], respectively.
\subsection{C \& O}
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f5.pdf}}
\caption{
[C/Fe] (left) and [O/Fe] (right) ratio for stars in Hypatia as a
function of [Fe/H], with a representative observational error bar in
the upper right. Each stellar abundance datapoint is colored according to the
radial distance of the host-star.
The median (blue) and 1$\sigma$ (between 16-84\%) quantile regression trends (black) are overlaid to better visualize the evolution of [C/Fe] with increasing [Fe/H]. The median and percentile values at [Fe/H] = -0.4, 0.0, and 0.4 dex are located in Table \ref{tab.sigma}.
}
\label{co}
\end{figure}
The element carbon, specifically the $^{12}$C isotope, is
formed via hydrostatic helium burning in stars, where the overall
production is governed by the competition between the
triple-$\alpha$ rate and destruction by the $^{12}$C($\alpha,\gamma$)$^{16}$O rate
\citep{Iben:1991p3494,wallerstein_1997_aa,busso_1999_aa,
Langanke:2007p3478,jose_2011_aa,bennett_2012_aa}.
Evolution of [C/Fe] as a function of [Fe/H] for the 808 stars in
the analysis of Hypatia is shown in Fig. \ref{co} (left).
A solar and relatively flat [C/Fe] ratio is interesting
because two competing sources come into play: intermediate and low mass stars begin depositing large amounts of
carbon but no iron, while SN Ia start injecting significant amounts of iron but no carbon.
To better understand the general trend of the data, the x-axis has been divided
into bins the size of the [Fe/H] representative errorbar, or 0.05 dex. In each bin, the median (blue) was determined as well as the 16th and 84th percentile of the data (black) using quantile regression. In this way the majority of the data, or 1 standard deviation ($\sigma$) pr 68\% between the percentile trend lines, highlights how [C/Fe] shifts with varying [Fe/H].
The blue and black curves in Fig. \ref{co} (left)
indicate a sharp decrease in [C/Fe] for [Fe/H] $<$ 0.0 dex, followed by a shallower decline with increasing [Fe/H]. This suggests that SN Ia injected more iron than the intermediate and low mass stars injected carbon at an earlier epoch.
There are $\sim$200 giant-class stars in the {\it Hypatia Catalog}, many of which were from the \citet{Thevenin:1998p1499} dataset and did not have carbon measurements. Only $\sim$50 out of the total 808 stars with carbon abundances were giants, where 25 of those stars had [C/Fe] $<$ 0.0 dex.
From Table \ref{tab.sigma}, we see that the 16th and 84th percentiles vary from the median by 0.11 and 0.16 dex respectively at [Fe/H] = 0.0 dex. As [Fe/H] increases to above-solar, where the majority of the data is located, the scatter of the 1$\sigma$ percentiles shrinks to 0.09 and 0.08 dex, respectively, at [Fe/H] = 0.4.
There also appears to be a slight concentration of
stars with high [Fe/H] but low [C/Fe] content at distances greater than 60 pc
-- see the last paragraph in this section for further analysis with oxygen.
\citet{Laird:1985p1923} determined carbon abundances in dwarf stars
using intermediate resolution ($\Delta \lambda$ = 1 \AA) image tube
spectra of the 3300-5250 \AA \ band features of molecular CH.
Many of these stars are in the analysis of the {\it Hypatia Catalog}, see \S \ref{s.struct}. Effective
temperatures were found from calibrated R-I, {\it b-y} and V-K color
indices. Surface gravities were derived from the spectra and
Str\"omgren photometry, supplemented with gravities based on parallax
data and estimated masses. A differential analysis was adopted, and
equivalent widths of the Fe I lines were used to determine the iron
abundances. Since no individual CH lines could be detected in
the spectra, LTE synthetic spectra
determined the final abundances. An analysis of the [C/Fe] ratio as a
function of the effective temperature indicated a systematic offset,
so a correction factor of 0.10 dex was applied to all the [C/Fe]
ratio. This correction factor was not used
in the {\it Hypatia Catalog},
but it provides discouraging insight into the effect of variations between datasets, specifically for carbon.
Oxygen is a product of hydrostatic He, C, and Ne burning, with $^{16}$O
being the dominant isotope \citep{Clayton:1968p4549,Arnett:1996p4446,
Thielemann:2002p3625,ekstrom_2011_aa}.
Three different oxygen features in the visible spectrum are used to determine
oxygen abundances: the O I triplet at 7700\AA\, , the [O I] doublet, or
the OH lines. Oxygen abundances determined from the excitation
feature (9.15 eV) O I triplet at 7700\AA \ are known to be sensitive
to the temperature structure of the model atmosphere, as well as being
affected by non-LTE corrections and convective inhomogeneities.
Nevertheless, all 933 stars in the analysis of Hypatia for which [O/Fe] was
determined use the O I triplet. Most catalogs applied various
empirical corrections by undertaking non-LTE
calculations or providing an agreement with
[O I] doublet determined abundances \citep[e.g.,][]{Edvardsson:1993p2124,
Brugamyer:2011p3104}.
Fig. \ref{co} (right) shows [O/Fe] versus [Fe/H] for the stars in the analysis of Hypatia.
The overall trend, shown by the blue and black trend lines, is classic $\alpha$-element,
starting from core-collapse supernovae depositing
large amounts of oxygen but no iron and later on SN Ia injecting
significant amounts of iron but no oxygen
\citep{gratton_1986_aa,marcolini_2009_aa,kobayashi_2011_aa}.
The 1$\sigma$ scatter in [O/Fe] is at least $\sim$ 0.25 dex for the entire range of [Fe/H], increasing to $\sim$ 0.40 dex ear the solar value of iron, or [Fe/H] = 0.0 dex (see Table. \ref{tab.sigma}). The inclusion of the \citet{Brewer:2006p1310} dataset had a direct impact on the expansion of the scatter in [O/Fe] for near-solar values of [Fe/H], although only by $\approx$ 0.1 dex, which was on par with error.
A large number of stars with [Fe/H] $>$ 0.2 dex also have a radial distance greater than 60pc from the Sun,
similar to the trend seen in the [C/Fe] plot. We investigated this further and determined that the high [Fe/H],
low [C/Fe] or [O/Fe] abundances at large radial distances is due to
the inclusion of \citet{petigura_2011_aa}. Their dataset provided the largest incorporation of both carbon and oxygen
abundances in 914 stars at an average distance of 52 pc (maximum 202 pc) from the Sun. Without their survey, stellar abundances exist in the high-iron, low-carbon/oxygen regime following the same trend, but in smaller numbers. While their data does not change the general features of the plots, it does smooth the progression for [Fe/H] $>$ 0.2 by adding more stars, specifically those at a distance beyond 60 pc. The preference for stars at high [Fe/H] and large distances may be a result of 1) survey bias, since the majority of the stars measured by \citet{petigura_2011_aa} were at [Fe/H] $>$ 0.0 dex (their Fig. 14) or 2) physical inhomogeneities in the Galactic disk, such as those discussed in \S \ref{gaps}; however more abundances at subsolar metallicities are needed before definite conclusions can be draw. The influence of the \citet{petigura_2011_aa} dataset is clear for the [C/Fe] and [O/Fe] abundances, but their data serves to solidify the trend observed by others. The addition of \citet{Ramirez:2007p1819} into Hypatia fulfilled a similar role for the [O/Fe] abundances, in that their abundances covered much of [Fe/H] $<$ 0.0 dex, which suffered from small number statistics. Their data, too, did not drastically change the overall feature of the [O/Fe] vs. [Fe/H] trend, but allowed for more reliable statistics.
\subsection{Mg,\,Si,\,S}
Magnesium is an $\alpha$-element whose dominant isotope $^{24}$Mg is
formed during hydrostatic carbon burning when a $^{12}$C+$^{12}$C
reaction creates the seed for $^{23}$Na($p,\gamma$)$^{24}$Mg, and
during hydrostatic neon burning via $^{20}$Ne($\alpha,\gamma$)$^{24}$Mg
\citep{Limongi:2003p3406,Karakas:2006p4657}.
Fig. \ref{mg} shows [Mg/Fe] as a function of [Fe/H] after the
renormalization to \citet{Lodders:2009p3091} (left) and according to the individual datasets (right).
Similar to the other $\alpha$-elements, there is a general decrease in the
[Mg/Fe] abundance in Fig. \ref{mg} (left) as [Fe/H] increases, due to
the late injection of iron from SN Ia, and a flattening of [Si/Fe]
at super-solar metallicities
\citep{matteucci_1986_aa,
Gibson:2003p2583,romano_2010_aa,kobayashi_2011_aa}.
However, the slope of [Mg/Fe] with [Fe/H] is shallower than the other two
$\alpha$-elements already examined, carbon and oxygen. The average 1$\sigma$ scatter is $\sim$ 0.2 dex in [Mg/Fe], becoming slightly larger for [Fe/H] = 0.0 dex and smaller at either extrema (see Table \ref{tab.sigma}). This variation in [Fe/H] supports the multiple productions sites predictions of \citet{Fenner:2003p4592}.
In Fig. \ref{co} (left), the stars at a further radial distance from the Sun had higher [Fe/H] ratios but lower [C/Fe];
the opposite appears to be true with respect to [Mg/Fe] such that stars at a distance of 60pc or less aren't as
enriched in magnesium as iron. In comparison, the majority of stars at 60pc or more have both [Mg/Fe] $>$ 0.0 dex and
[Fe/H] $<$ 0.0 dex. The majority of nearby stars exhibit super-solar [Fe/H] abundance ratios. Overall, these trends hint at the possibility of there existing two ``ensembles" of stars: one with high-[Mg/Fe] but low [Fe/H] at larger distances from the Sun and one with the opposite, separated by a ``gap" around [Mg/Fe] $\approx$ -0.05 dex. \citet{Edvardsson:1993p2124,Venn:2004p1483,Mishenina:2008p1380} all show [Mg/Fe] vs. [Fe/H] for disk-stars with a small ``gap" separating the two groups that corroborate our trend. We compared the impact of the inclusion of many large datasets in Hypatia by temporarily excluding them and then reanalyzing the abundances. Specially, we temporarily omitted any catalog listed in Table \ref{tab.cat} that had 200 stars or more by removing each dataset individually, recompiled the Hypatia Catalog, and then produced new plots. A comparison of the new plots-minus-one-catalog with respect to the full catalog plots gave a good indication on each dataset's overall influence. No catalogs caused a dramatic change in the [Mg/Fe] trend which could explain the two ``ensembles" or ``gap." See \S \ref{gaps} for more discussion.
An obvious effect from the renormalization was a change in the [Mg/Fe] vs. [Fe/H] slope between the left and right plots in Fig. \ref{mg}. The individual datasets used a variety of different solar abundances (see Table \ref{tab.long}), such that $\log \epsilon(Mg)$ had a range of 7.43--7.77.
Since $\log \epsilon(Mg)$ = 7.54 according to \citet{Lodders:2009p3091}, which fell within the range of used of solar-Mg normalizations, the re-normalization affected each of the [Mg/Fe] abundances differently.
Looking at the 1$\sigma$ quantile regression lines between the two plots in Fig. \ref{mg}, it is clear that the scatter is lower when using the solar abundances given by the individual datasets (see Table \ref{tab.sigma} for the mean and 1$\sigma$ values for the renormalized data). However, the absolute difference between the un-normalized ($A_{un}$) and re-normalized ($A_{re}$) abundances, or $| A_{un} - A_{re} | $, for all element ratios ([X/Fe]) in the {\it Hypatia Catalog} has a mean of 0.06 dex and a median of 0.04 dex. In other words, to correct for varying solar abundance scales, the renormalization resulted in an average 0.06 dex shift in the abundance measurements. Specifically for Fig. \ref{mg}, the mean and median absolute differanaences for [Mg/Fe] = 0.15 and 0.12 dex, respectively, while the representative error is 0.07 dex. This variation is on par with representative error bar for most elements, especially those that are not neutron-capture, and has not been accounted for during survey comparisons. While the renormalization corrections create a drastic change in the data between the two plots in Fig. \ref{mg}, these adjustments need to be adopted (amongst others) in order to uncover the physical trends in the data. Fig. \ref{mg} (right), without the solar renormalization, hides the two ``ensembles" we noted to the left and discuss in \S \ref{gaps}, and does not accurately reflect the scaling variations between datasets (see Table \ref{tab.long}).
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f6.pdf}}
\caption{
Similar to Fig. \ref{co} but for magnesium, showing the
data after being re-normalized (left) to \citet{Lodders:2009p3091} and
as measured by the literature sources (right).
}\label{mg}
\end{figure}
The dominant $^{28}$Si isotope is produced by hydrostatic and
explosive oxygen burning in massive stars \citep{Arnett:1996p4446}.
Fig. \ref{sis} (left) shows silicon is a classic $\alpha$-chain element
such that [Si/Fe] decreases with increased [Fe/H].
The [Si/Fe] ratio has the most entries in the {\it Hypatia Catalog}, being
measured for 1098 thin-disk stars with well agreed upon literature determinations.
There is a 1$\sigma$ scatter $\approx$ 0.18 dex about the median trend (black line) at any given [Fe/H], which is at a maximum around solar, 0.23 dex (see Table \ref{tab.sigma}).
Similar to [C/Fe] in Fig. \ref{co} (left) and [Mg/Fe] in Fig. \ref{mg} (left),
there is a slight gap in abundances for [Si/Fe] around $\approx$ -0.1 dex. However, this ``gap" is
sub-solar as opposed to [C/Fe] and the second, lower ``ensemble" of stars with higher [Fe/H]
seem to be comprised of stars that have a radial distance of 60 pc or less, similar to [Mg/Fe] in Fig. \ref{mg} (left).
For all of the stars in the lower ``ensemble," about half of them ($\sim$ 290) originated from the \citet{Valenti:2005p1491} catalog. The remaining $\sim$ 240 stars were contributed by other datasets, such a \citet{Fulbright:2000p2188, Gilli:2006p2191, Takeda:2007p3681, Mishenina:2008p1380, Neves:2009p1804, DelgadoMena10}. As a test regarding the influence of the \citet{Valenti:2005p1491} dataset, we recompiled the {\it Hypatia Catalog} without the inclusion of their measurements and found that the ``gap" was still present at [Si/Fe] $\approx$ -0.1 dex for all [Fe/H].
Rather than cover up the ``gap," the inclusion of \citet{Valenti:2005p1491} reinforced the trend delineated by the other datasets, seen in Fig. \ref{co} (left).
The differences between these two ``ensembles" indicate that there may be abundance correlations with distance and that the local neighborhood is not homogeneously mixed (see \S \ref{gaps} for more discussion).
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f7.pdf}}
\caption{
Same as Fig. \ref{co} but for silicon (left) and sulfur (right).
[Si/Fe] is one of the most common measurements in the analysis of Hypatia (1098 stars),
with less entries for [S/Fe] (162 stars).
Due to small number statistics, the median and percentile trend lines could not be accurately
determined for [S/Fe], which has less than 250 stellar measurements (see Table \ref{tab.skip}).
}\label{sis}
\end{figure}
Sulfur is produced within massive stars, via hydrostatic and explosive oxygen
and burning \citep{Clayton:1968p4549,heger_2000_aa,
rauscher_2002_aa,Limongi:2003p3406}.
There are relatively fewer stars, 162 in the analysis of Hypatia, for which sulfur
has been measured, as shown in Fig. \ref{sis} (right). This is due to
absorption lines being too weak in the visible spectrum or blended to
separate from the continuum, making it difficult to determine an
accurate abundance \citep{Francois:1987p2332}. \citet{takeda_2005_aa} reports significant, $\approx$ 0.2 dex,
non-LTE corrections affecting several lines used in the determination
of S and Zn abundances in F, G, and K stars.
The small number of stars in Fig. \ref{sis} (right) made the calculation of the median and 16/84\% percentile regression lines impossible.
Like other $\alpha$-elements, there is decrease in [S/Fe] as [Fe/H] increases from $\approx$ -0.6
dex to $\approx$ 0.6 dex and a slope similar to [Si/Fe] (left). As previously noted
for [Si/Fe] vs. [Fe/H], [S/Fe] exhibits a gap in element abundance between [0.0, 0.1] dex (see \S \ref{gaps} for more discussion).
The scatter in [S/Fe] is $\approx$ 0.3 dex over the entire [Fe/H] range shown, but the two ``ensembles" makes
it hard to estimate. \citet{Mishenina:2008p1380} also show two ensembles when analyzing the [Si/Fe] vs. [Fe/H] abundance trends.
\citet{Luck:2005p1439} reported the abundances for Mg, Si, and S, as
well as for 25 other elements, in 114 F, G, K, M stars within 15 pc of
the Sun. Table \ref{tab.cat} shows 110 of these stars are in the {\it
Hypatia Catalog}. Their high signal-to-noise spectra (in excess of
$\approx$ 150 per spectral pixel) were taken between 1997 and 2003
using the Sandiford Cassegrain Eschelle Spectrograph attached to the
2.1 m telescope at McDonald Observatory. They determined the solar
flux spectrum by using differential analysis, with Callisto as the
reflector. The model atmospheres were determined by MARCS75
\citep{Gustafsson:1975p4658}. Photometry was acquired through the
General Catalogue of Photometric Data \citep{Hauck:1991p4695}.
Surface gravities $\log g$ values and Fe abundances were obtained by
iterating until the [Fe/H] value from both Fe I and Fe II were equal.
Overall abundance uncertainties for [X/Fe] were determined on a per
element basis.
\subsection{Ca \& Ti}
Calcium is an $\alpha$-element whose dominant, double magic isotope $^{40}$Ca
is produced by oxygen burning in massive stars \citep{Woosley:1995p3481}.
While most of the catalogs within the analysis of Hypatia
determined their calcium abundances through
Ca I, two catalogs \citep{AllendePrieto:2004p476, Gebran:2010p6243}
used Ca II lines, shown in Fig. \ref{ca} (right).
\citet{AllendePrieto:2004p476} compared the derived abundances
from the neutral and ionized lines, and reported the abundance
for Ca I and II differed by 0.25 dex due to the broadening of the
wings in the line profiles. These dissimilarities could be mollified
by a change in the surface gravity and T$_{eff}$, at the expense of
weakening the Ca I line. Therefore, their final abundances were
derived from the Ca II 8662 \AA\ line, since this was less blended
than the Ca II 8498 \AA\ line.
Fig. \ref{ca} shows how [Ca/Fe] exhibits the same trend with
[Fe/H] as other $\alpha$-elements. However, the shallow slope over the
[Fe/H] range suggests that calcium production by massive
stars is more closely balanced by iron production from SN Ia.
There
is $\approx$ 0.09-0.19 dex in the 1$\sigma$ scatter for [Ca/Fe] as [Fe/H] increases from -0.4 dex to 0.4 dex, per Table \ref{tab.sigma}.
The abundances for [Ca II/Fe] vs. [Fe/H] are not shown because the total number of stars,
after potential thick-disk stars and abundance measurements beyond error bar were removed,
totaled only 8 stars.
It also appears that stars that
are further away from the Sun, at a distance greater than 60 pc (red circles), exhibit generally
higher abundances of [Ca/Fe].
However, a closer investigation reveals that the majority of stars with [Ca/Fe] $<$ -0.1 dex were measured solely by \citet{Neves:2009p1804}. While they did not place a direct distance cut on the stars they analyzed, their data originates from the HARPS GTO (see Table \ref{tab.long}), which do not observe stars that are greater than 56 pc from the Sun. The abundances from \citet{Neves:2009p1804} provided [Ca/Fe] measurements from $\approx$ 400 unique (i.e. only measured by them) stars. Their methodology is consistent with others in the field (see Table \ref{tab.long}) such that their own comparison with literature sources did not show any significant differences. Without the inclusion of the \citet{Neves:2009p1804} survey, there are a number of stars in the region with high [Ca/Fe] and low [Fe/H], as measured by other surveys. A few of these stars are located at distances $>$ 60 pc from the Sun. As a result, the feature that separates the large-distance, high [Ca/Fe] and low [Fe/H] stars from their near-distance, low [Ca/Fe] and high [Fe/H] counterparts may not be physical. To be certain, we would either need to obtain more abundance measurements for stars a distances $>$ 60 pc from the Sun to obtain better statistics or reanalyze all of the stellar data from the surveys in the {\it Hypatia Catalog} with Ca I lines using a standardized method.
In a similar vein, we found that a number of stars that had a solar value of [Ca/Fe] but with [Fe/H] $>$ 0.2 dex originated from \citet{Trevisan:2011p6253}. Further study found that \citet{Trevisan:2011p6253} contributed a unique and small outcrop of stars at high [Fe/H] but near solar-value of [X/Fe] for three elements: Ca, Ti, and Ni. Their Figure 21 confirms these variations with respect to other catalogs, namely \citet{Bensby:2003p513, Bensby:2004p529, Mishenina:2004p1360, Mishenina:2008p1380, Reddy:2003p1354, Reddy:2006p1770}. They also report an average difference in abundance determinations within $\sim$ 0.2 dex compared to \citet{Neves:2009p1804, Valenti:2005p1491}, which is much greater than standard error for Ca, Ti, and Ni. This is unlike other cases where features in the abundance distribution may be dominated by a single survey, but are still present with reduced numbers of stars when the survey is removed, like the ``gaps" in abundances (see \S \ref{gaps}).
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f8.pdf}}
\caption{
Same as Fig. \ref{co} but for calcium.
}\label{ca}
\end{figure}
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f9.pdf}}
\caption{
Same as Fig. \ref{co} but for neutral (left) and ionized (right)
titanium.
Due to small number statistics, the percentile trend lines could not be accurately
determined for [TiII/Fe], which has less than 250 stellar measurements (see Table \ref{tab.skip}).
}\label{ti}
\end{figure}
Titanium is produced in massive stars by
explosive burning processes in core-collapse supernovae
\citep{woosley_1973_aa,Arnett:1996p4446,Limongi:2003p3406}.
Like calcium, titanium abundances have been determined with two
ionization states as shown in Fig. \ref{ti}. Since there are a large
number of spectral lines for Ti I and Ti II in the optical spectrum,
titanium is one the more commonly measured elements in Hypatia
(see Fig. \ref{fig.hist}).
While a number of catalogs measured abundances using both the Ti I and Ti II lines
\citep{Bond:2008p2099,Gratton:2003p1182,Neves:2009p1804,Takeda:2007p1531},
\citet{Bergemann:2011p4775} found that abundances
can vary by 0.1 dex or larger when comparing pure Ti I and Ti II line determinations.
Most catalogs, though, used the Ti I lines alone.
The [Ti/Fe] binned trend with [Fe/H], shown by the trend lines in
Fig. \ref{ti} (left), suggests an evolution similar to the other $\alpha$-elements.
Although, it
should be noted that the dominant isotope is produced by alpha chain nucleosynthesis, being a beta decay product of the alpha isotope $^{48}$Cr.
There is a $\approx$ 0.15
dex 1$\sigma$ scatter in [Ti/Fe] over the entire range of [Fe/H] in Fig. \ref{ti} (left).
For [Fe/H] $<$ 0.0 dex, stars at larger distances (red) tend to show near-solar [Ti/Fe],
while stars at smaller distances tend to exhibit lower [Ti/Fe] and higher [Fe/H].
The \citet{Valenti:2005p1491} catalog contributed the majority of stars with [Fe/H] $>$ 0.1 dex, or $\sim$ 80\% of their stars with titanium abundances. Out of those with enriched iron, only 8\% were at a distance greater than 60 pc from the Sun, where the average was 36 pc. The distance-abundance trend seen for [Ti/Fe] may therefore be a product of target selection and is left for further investigation. However, the steady decline of [Ti/Fe] with increasing [Fe/H] was confirmed by other surveys who measured stars in this region, but with much a much smaller sample of stars. Similarly, \citet{Trevisan:2011p6253} contributed the majority of stars with [Fe/H] $>$ 0.2 dex and [Ti/Fe] $>$ 0.0 dex, as previously discussed, but was not the lone source for stars in this region.
The median trend for [Ti II/Fe] is similar to [Ti/Fe],
Fig. \ref{ti} (right),
however with fewer stars such that the 1$\sigma$ trend lines cannot be accurately determined.
The median trend indicates that the slope
of [TiII/Fe] becomes more shallow as [Fe/H] increases, which may be a
result of smaller-number statistics. There are significantly fewer
stars with [Ti II/Fe] determinations at distances greater than 60 pc. Both \citet{Venn:2004p1483,Mishenina:2008p1380} saw
separate ensembles in their abundance determinations for [Ti/Fe]. However, possibly due to the separation of neutral and ionized titanium in our analysis, we do not reproduce their results.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f10.pdf}}
\caption{
The same [Ti/Fe] abundances as given in Fig. \ref{ti} but with respect to distance,
where the abundance trends are shown with respect to height above the Galactic plane (left) and direction
to the Galactic center and anti-center (right). The error bars are similar to Fig. \ref{ferad}.
}\label{tirad}
\end{figure}
Fig. \ref{tirad} (left) shows [Ti/Fe], an $\alpha$-element, as a function of distance with
error bars similar to Fig. \ref{ferad}.
On the left, the stars are colored according to their height above the
Galactic plane, $z$.
The [Ti/Fe] abundances show little dependence on height above the plane, where they appear
evenly mixed at all radial distances, although fewer stars near the mid-plane were measured
at further distances (red circles).
And the majority of stars in the analysis of Hypatia that have [Ti/Fe]
abundance measurements are located within 60 pc of the Sun.
Fig. \ref{tirad} (right) shows those same stars, but colored according
to their position toward the galactic center.
Here as well, there was no discernible trend with distance, although fewer stars located towards the
Galactic center were measured at further radial distances.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f11.pdf}}
\caption{
Similar to Fig. \ref{co} but for [Na/Ca] (left) and [Si/Ca] (right) as a function of [Fe/H].
}\label{naalpha}
\end{figure}
Fig. \ref{naalpha} shows [Na/Ca] (left) and [Si/Ca] (right)
as a function of [Fe/H].
The [Na/Ca] evolution, an odd-Z element to an $\alpha$-element ratio,
shows a change in slope of [Na/Ca] at [Fe/H] $\approx$ 0.0 dex
\citep{marcolini_2009_aa}.
Production of sodium and calcium were roughly
equivalent at smaller [Fe/H], but sodium dominates
calcium as [Fe/H] increases,
with a relatively constant $\approx$ 0.2 dex 1$\sigma$ scatter in [Na/Ca] for all [Fe/H].
This trend may be due to SN Ia or intermediate- to low-mass stars
injecting additional sodium relative to calcium at later times.
The outlying stars with super-solar [Na/Ca] and [Fe/H] $>$ -0.2 may be due to late contributions form asymptotic giant branch (AGB) stars.
In contrast, [Si/Ca] with [Fe/H] (Fig. \ref{naalpha}, right) shows a flat and solar trend within error.
The $\approx$ 0.2--0.1 dex 1$\sigma$ scatter shows how silicon and calcium, both $\alpha$-elements, are
dominated by contributions from massive stars which co-produce both Si and Ca. The similar nucleosynthetic origin sites, along with the relative ease of measuring all three elements (Na, Si, and Ca) within stars (see Fig. \ref{fig.hist}), may explain the small scatter for these abundance ratios. Other implied subtleties may warrant future consideration.
\section{Odd-Z Elements (N, Na, Al, K, \& Sc)}
\label{s.odd}
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f12.pdf}}
\caption{
[N/Fe] as a function of [Fe/H] with the same format as Fig. \ref{co}.
Due to small number statistics, the median and percentile trend lines could not be accurately
determined for [N/Fe], which has less than 250 stellar measurements (see Table \ref{tab.skip}).
}
\label{lin}
\end{figure}
Isotopes of nitrogen are produced in stars by the CNO cycle
\citep{Arnett:1996p4446}, where primary nitrogen is usually
produced as a convective helium burning shell mixes into a
hydrogen shell, where C and O nuclei form nitrogen with
nearly explosive consequences
\citep{Talbot:1974p4584,meynet_2002_aa,
ekstrom_2008_aa,karakas_2010_aa}.
Fig. \ref{lin} shows [N/Fe] with respect to [Fe/H],
where the axes for all the odd-Z elements are now [Fe/H] = [-0.65, 0.65] and [X/Fe] = [-0.6, 0.6].
There are $\sim$15 times fewer stars for which [N/Fe]
has been measured as compared to the [C/Fe] ratio in our analysis of the {\it Hypatia
Catalog} (see \S \ref{s.struct} and Table \ref{tab.skip}), making it a priority measurement for future observations.
As a result, neither the median nor the 1$\sigma$ regression lines could be determined.
The majority of [N/Fe] abundances are above solar for [Fe/H] $<$ 0.2 dex, with few measurements for stars with [Fe/H] above solar. The scatter in [N/Fe] is rather significant for all [Fe/H] abundances. If physical, the large scatter suggests $^{14}$N was produced as a primary element, since [N/Fe] is relatively constant with [Fe/H]
\citep{Laird:1985p1923,carbon_1987_aa}. However, given the difficulty in measuring [N/Fe], the large scatter may be due to discrepancies between data reduction techniques.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f13.pdf}}
\caption{
The [Na/Fe] ratio as a function of [Fe/H]. The figure has the same format
as Fig. \ref{mg} where the figure on the
left shows [Na/Fe] after the renormalization to \citet{Lodders:2009p3091} and the
abundances on the right use the solar abundances determined by the literature sources.
}
\label{na}
\end{figure}
The only stable isotope of sodium, $^{23}$Na, is produced mainly in
carbon-burning in massive stars, whose final abundance is sensitive to
the overall neutron enrichment \citep{Woosley:1995p3481,
Chieffi:2004p5067}.
The abundance ratio [Na/Fe], after being renormalized
to a standard solar abundance (see \S \ref{s.spread}), as a function of the [Fe/H] ratio is shown
in Fig. \ref{na} (left) for the 907 stars in the analysis of the {\it Hypatia Catalog}, see \S \ref{s.struct}.
The median and 1$\sigma$ lines shows a shallow decreasing trend that curves up at higher [Fe/H] metallicities, similar to that seen in \citet{Edvardsson:1993p2124,Bensby:2003p513,Bensby:2005p526}. The minimum of the median trend occurs below solar at [Na/Fe] $\approx$ -0.2 dex and [Fe/H] $\approx$ 0.2 dex, such that most of the stars observed have sub-solar [Na/Fe] abundances. The majority of stellar abundances for [Fe/H] $>$ 0.2 dex came from both \citet{Neves:2009p1804, Valenti:2005p1491}, reinforcing the positive slope in this region seen in smaller numbers from other surveys. Overall, the 1$\sigma$ scatter increases from 0.13 dex at [Fe/H] = -0.4 dex to 0.19 dex at [Fe/H] = 0.4 dex, see Table \ref{tab.sigma}. However, there is an increased amount of scatter in [Na/Fe] for [Fe/H] = [0.0, 0.2] dex, note the 1$\sigma$ percentile is 0.26 at [Fe/H] = 0.0 dex per Table \ref{tab.sigma}. Those stars with [Na/Fe] $>$ 0.2 dex are predominantly from \citet{Thevenin:1998p1499,Gebran:2010p6243}, with a few contributions from \citet{Luck:2006p6883,Valenti:2005p1491,Edvardsson:1993p2124,Takeda:2007p3681,Reddy:2006p1770,Feltzing:1998p886}.
Fig. \ref{na} (right)
shows the effects of the individual abundance scales applied by each of the datasets. While there are
some changes to the scatter between the two figures (left and right), the more noticeable distinction
is the how the overall trend in [Na/Fe] to [Fe/H] to the left is more scattered compared to the right (see Table \ref{tab.sigma}).
As discussed for Fig. \ref{mg}, the variation in abundance slopes in Fig. \ref{na} indicates that the solar abundance scales are an important factor when comparing datasets (see Table \ref{tab.long}). The significant shifting, where the mean and median absolute difference between the un-normalized and re-normalized [Na/Fe] abundances were 0.18 and 0.16 dex, respectively, causes stars to reposition themselves in new regions of the plot, which then varies the underlying trends. Using the same solar abundance scale is only one step out of many that need to be taken to accurately compare datasets between groups.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f14.pdf}}
\caption{
Similar to Fig. \ref{co} but for aluminum.
}\label{al}
\end{figure}
Aluminum, whose only stable isotope is $^{27}$Al,
is mainly synthesized in hydrostatic carbon and neon burning
\citep{Arnett:1985p4777,Thielemann:1985p4797,Woosley:1995p3481,limongi_2006_aa}.
Evolution of [Al/Fe] with [Fe/H] is shown in Fig. \ref{al} for the 523 stars in the
analysis of the {\it Hypatia Catalog}, see \S \ref{s.struct}.
The three trend lines suggest [Al/Fe] values are near-or-below solar with a shallow decreasing trend, consistent with the trends seen in data studies over a larger
metallicity range \citep{Peterson:1981p4862, Magain:1989p4878,
Fulbright:2000p2188,Brewer:2006p1310}.
For the majority of [Fe/H], the 1$\sigma$ scatter in [Al/Fe] is relatively constant (see Table \ref{tab.sigma}). Although, curiously, there is a relative paucity of [Al/Fe] measurements for stars with [Fe/H] $\gtaprx$ 0.4 dex. The few outlier stars with [Al/Fe] above 0.2 dex are from a wide range of catalogs.
Fig. \ref{alrad} shows [Al/Fe] with respect to radial distance,
colored to show height above the Galactic plane (left) and directionality
towards or away from the galactic center (right). Stars within 60 pc span a slightly larger
range of [Al/Fe], -0.3 $\ltaprx$ [Al/Fe] $\ltaprx$ 0.4, than stars
at larger distances. But this variation is on par with the error bars associated with [Al/Fe]. Using aluminum as a typical odd-Z element, there does not appear
to be any distinct correlation between abundance and location in the local galaxy.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f15.pdf}}
\caption{
[Al/Fe] ratio as a function of radial distance, similar to to
Fig. \ref{tirad}.
}
\label{alrad}
\end{figure}
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f16.pdf}}
\caption{
Similar to Fig. \ref{co} but for potassium.
Due to small number statistics, the percentile trend lines could not be accurately
determined for [K/Fe], which has less than 250 stellar measurements (see Table \ref{tab.skip}).
}
\label{pk}
\end{figure}
The isotopes
$^{39,41}$K are produced during oxygen burning, with $^{41}$K made as
$^{41}$Ca, but some $^{41}$K is produced as itself during neon
burning.
The abundance ratio [K/Fe] versus the [Fe/H] ratio is shown in
Fig. \ref{pk} for the 139 stars in the analysis of the {\it Hypatia Catalog}, see \S \ref{s.struct}. Blending
between the K I and atmospheric O$_2$ lines means potassium abundances
are more difficult to determine \citep{gratton_1987_aa}, accounting
for the relatively fewer number of stars with K abundance determinations.
In addition, the low excitation energies of K I might be susceptible
to non-LTE or strong hyperfine structure effects
\citep{ivanova_2000_aa}.
Despite the median blue line in Fig. \ref{pk}, the large scatter in [K/Fe] makes it difficult
to identify any sort of general trend. The only point of interest from Fig. \ref{pk} is the
complete lack of stars with [K/Fe] abundances at [Fe/H] $>$ 0.11, despite the 6 data sets within
Hypatia that have measured potassium (see Table \ref{tab.cat}).
A number of abundance determinations from \citet{Reddy:2003p1354} are used in
Fig. \ref{na} (right) and Fig. \ref{pk}. They investigated 27
elements, including sodium and potassium, in 181 F and G dwarfs from a
differential LTE analysis of high-resolution ($\Delta \lambda /
\lambda$ $\approx$ 60,000) and high signal-to-noise (S/N=300-400)
spectra from the Smith 2.7 m telescope at McDonald Observatory. Of
these 181 stars, 179 are in Hypatia. Effective temperatures were
adopted from an infrared flux calibration of St\"romgren
photometry. Surface gravities and stellar ages were determined from
stellar evolution tracks and {\it Hipparcos Catalogue} parallaxes.
The 6154.23 \AA \ and 6160.75 \AA \ lines of Na I and the 7698.98 \AA
\ line of K I were used in the analysis, with oscillator strengths
taken from \citet{Lambert:1968p5112}.
The isotope $^{45}$Sc is made as itself and as
radioactive $^{45}$T \citep{Tuli:2005p6038}
in hydrostatic and explosive oxygen-burning and in alpha-rich
freezeouts in core-collapse events \citep{rauscher_2002_aa,
Limongi:2003p3406,ekstrom_2011_aa}.
The trend of [Sc/Fe] versus [Fe/H] is shown in Fig. \ref{sc} for the
381 stars with Sc I measurements (left) in the analysis of the {\it Hypatia Catalog}, see \S \ref{s.struct}, and the 386 stars with Sc II abundance determinations (right).
The [Sc/Fe] ratios are relatively flat with respect to [Fe/H],
where the median of the abundance measurements is slightly above solar at low [Fe/H]
ratios and slightly below solar at higher metallicities (see Table \ref{tab.sigma}). The 1$\sigma$
trend lines point out the small number of stars with [Sc/Fe] measurements for [Fe/H] $<$ -0.2 and $>$ 0.4.
Their scatter around the solar iron ratio is 0.28 dex.
In contrast, [Sc II/Fe] ratios follow a more steeply decreasing trend with [Fe/H]
\citep{Thevenin:1998p1499,Zhang:2006p6230}.
The small 1$\sigma$ scatter at either end of the [Fe/H] extrema, 0.14 and 0.09 dex, respectively, is mitigated by the relatively large 0.23 dex scatter at the solar value of [Fe/H]. While a few stars of these stars also have [ScII/Fe] above 0.3 dex, their measurements come from a number of different catalogs. The majority of the scatter may be dependent on whether individual studies correctly included hyperfine structure in their Sc II abundance determinations. To this end, we found that only 3 of the 15 catalogs, \citet{Feltzing:1998p886,AllendePrieto:2004p476,Gebran:2010p6243}, that measured Sc II ratios incorporated hyperfine structure in their analysis. Analyzing the [ScII/Fe] abundances from only these three catalog still produces some scatter. However, given that the total number of stars was 159, we don't find our results to be conclusive.
Similar to the trends seen for [Mg/Fe], [Si/Fe], and [S/Fe] (see Figs. \ref{co}--
\ref{sis}), the stars for which [ScII/Fe] was measured show two ``ensembles"
separated by a ``gap" between [ScII/Fe] $\approx$ [0.0, 0.1].
Further investigation revealed that the majority of the lower ``ensemble" was a result of the \citet{Neves:2009p1804} dataset. The inclusion of their catalog increased the number of stars with [ScII/Fe] $<$ -0.2 by an order of magnitude. The contribution of the additional, unique stars acted to solidify the trend that was already present from other datasets, although in smaller numbers. We therefore maintain that this ``gap" is physical, as opposed to a signature of catalog compilation. In a similar vein as the discussion for [Ca/Fe] in Fig. \ref{ca}, we find that the predominant number of stars near to the Sun in the lower ``ensemble" is a direct consequence of the \citet{Neves:2009p1804} dataset only observing stars at distances $<$ 60 pc. For a more complete analysis, we would require additional abundance measurements for stars more than 60 pc from the Sun or a standardized method for determining stellar abundances. We discuss the two ``ensembles" seen in the [ScII/Fe] vs. [Fe/H] plot in \S \ref{gaps}.
\citet{Feltzing:1998p886} explored scandium abundances in 47 G and K
dwarf stars with -0.1 dex $<$ [Fe/H] $<$ 0.42 dex using a differential
LTE analysis with respect to the Sun of high-resolution ($\Delta
\lambda / \lambda$ $\approx$ 100,000) and high signal-to-noise (S/N
$\approx$ 200) spectra. Of these 47 stars, 45 are in Hypatia (see
Table \ref{tab.cat}). The 5484.64 \AA \ line is used for Sc I and the
5239.82 \AA \ 5318.36 \AA \ 6300.69 \AA \ 6320.84 \AA \ lines are used for Sc II.
Noting that single line abundance
determinations should be viewed with caution,
\citet{Feltzing:1998p886} base their scandium abundance determinations
on Sc II and discuss the apparent overionization and other non-LTE
effects that may effect most of the abundance determination, including
scandium. \citet{Zhang:2008p6102} performed a non-LTE study of
scandium in the Sun and find strong non-LTE abundance effects in Sc I
due to missing strong lines. Thus, scandium abundances determined from
single line LTE determinations are generally unsafe and abundances
based on multiple Sc II lines in non-LTE are preferred.
\vskip -1.0in
\begin{figure}[htb]
\centering
\centerline{\includegraphics[height=2.5in]{f17.pdf}}
\caption{
The [Sc/Fe] ratio as a function of [Fe/H], with the same
format as Fig. \ref{co}.
}
\label{sc}
\end{figure}
\FloatBarrier
\section{Iron-Peak Elements (V, Cr, Mn, Co, \& Ni)}
\label{s.ironpeak}
The elements V, Cr, Mn, Co, and Ni are formed by the same nuclear
processes that create iron in core-collapse and thermonuclear supernova, in varying degrees
\citep{Thielemann:2002p3625,Limongi:2003p3406,Thielemann:2007p4473}.
Vanadium is dominated by the isotope $^{51}$V, which is produced as
$^{51}$Cr and $^{51}$Mn during explosive oxygen burning, explosive
silicon burning, and $\alpha$-rich freezeouts in core-collapse
supernovae \citep{Clayton:2003p4494}.
Chromium, essentially $^{52}$Cr, is formed as a result
of radioactive decay from $^{52}$Fe during quasiequilibrium explosive
silicon burning \citep{Arnett:1996p4446,dauphas_2010_aa}.
Manganese -- dominated by
$^{55}$Mn from the radioactive decay of $^{55}$Co, cobalt -- dominated by
$^{59}$Co from the radioactive decay of $^{59}$Cu, and nickel --
dominated by $^{58}$Ni made as itself, are all generally the result of
quasiequilibrium reactions during explosive silicon burning
\citep{Woosley:1995p3481}. Because of these elements' proximity to iron (see \S
\ref{s.iron}), most of the abundance evolutions track iron.
However, given the noted scatter for the elements in this subsection, there is a striking case for improving the oscillator strength values for the iron-peak abundances. All of the iron-group elements have been plotted with the same x- and y-axis scales, where [Fe/H] = [-0.6, 0.6] and [X/Fe] = [-0.65, 0.5], respectively.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f18.pdf}}
\caption{
Same as Fig. \ref{co} but for vanadium.
}
\label{v}
\end{figure}
Of the 28 literature sources in the {\it Hypatia Catalog} that
determined vanadium abundances, only \citet{Feltzing:1998p886} and
\citet{Takeda:2007p3681} determined vanadium abundances from both
ionization states. Both surveys reported the lines for the neutral
and ionized species were limited to only one or two lines in the
optical spectrum, or too weak to separate out from the spectrum.
\citet{Zhang:2008p6102} reported the vanadium abundances using V I,
for 32 mildly metal poor stars using
spectra with a signal-noise ratio of about 150 per pixel at 6400 \AA\,
and a resolving power of about 37,000. Solar abundances, calculated
from the daylight spectrum were used to derive stellar abundances
relative to the Sun. The effective temperature was determined
from the {\it b-y} and {\it V-K} color indices; surface gravities were
calculated from Hipparcos parallax. They reported that V I follows Fe
very closely, with no offset between thin and thick disk stars.
The ratio [V/Fe] versus [Fe/H] is shown in Fig. \ref{v} for the 564 stars in the analysis of the {\it Hypatia Catalog}, see \S \ref{s.struct}.
The median [V/Fe], shown by the blue line,
indicates a flat and slightly sub-solar trend for all [Fe/H] content, which is consistent with its nucleosynthetic origin.
The average scatter, as determined by the 1$\sigma$ percentile lines, is relatively large: 0.22 dex (see Table \ref{tab.sigma}). This scatter was persistent across many individual datasets and may be the result of hyperfine structure being treated rather ``casually" for the V I lines.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f19.pdf}}
\caption{
Similar to Fig. \ref{tirad}, but for vanadium.
}\label{vrad}
\end{figure}
Fig. \ref{vrad} shows [V/Fe] ratios versus radial distance, colored to
show the height above the Galactic plane (left plot) and
directionality towards or away from the galactic center (right plot).
There is a large scatter of $\approx$ 0.8 dex in [V/Fe] for
those stars with a distance less than $\sim$ 40 pc and away from the galactic center (blue and green circles).
The large scatter at radial distances less than 40 pc is an interesting and unexpected result. The majority of abundances beyond 40 pc lie below [V/Fe] $=$ 0.1 dex, however near to the Sun [V/Fe] = [-0.3, 0.5]. In addition, stellar abundances greater than 0.1 dex come from a variety of different surveys. Because there are a number of unblended V I lines in the optical spectrum, we do not believe this cutoff is a function of survey or instrument bias. For now, we leave this question as something to examine more thoroughly in the future. To the right of Fig. \ref{vrad}, there were fewer stars for which [V/Fe] was measured near the galactic plane at large radial distances. However, it does appear that stars towards the Galactic anti-center (purple) tend to cluster around solar [V/Fe] at all distances.
Evolution of [Cr/Fe] with [Fe/H] is shown in Fig. \ref{cr} for the
483 stars in the analysis of {\it Hypatia Catalog} with both neutral Cr abundances
determinations (left) and the 329 stars with Cr II based abundance
determinations (right).
Thirty-one catalogs within Hypatia report
abundances from either pure Cr I lines or a blend with Cr II, while nine surveys published using only
Cr II lines (see Table \ref{tab.cat}).
For example, \citet{Neves:2009p1804} present a survey 12 elements whose
abundances are derived from spectra obtained with the HARPS
spectrograph on the ESO 3.6 m telescope. The Cr I lines 4588.20
\AA\ and 4592.05 \AA \, along with the Cr II line of 4884.61 \AA
\ were used in a differential LTE analysis relative to the Sun to
determine the abundance levels. Of the 451 stars in the
\citet{Neves:2009p1804} survey, 443 are in the {\it Hypatia Catalog}.
Initial estimates of the oscillator strengths were taken from the
Vienna Atomic Line Database and refined using a semi-empirical,
inverse analysis with the MOOG2002 \citep{Sneden:1973p6104}. Effective
temperatures, surface gravity, microturbulence, and metallicity were
taken from \citet{Sousa:2008p2748}. \citet{Neves:2009p1804} reported
that abundance levels determined from neutral states are more
sensitive to effective temperature changes, whereas abundances derived
from ionized states are more sensitive to changes in surface
gravity. Abundances from ionized elements are also more sensitive to
metallicity changes than the neutral elements, although the
sensitivity is not as significant as for the effective
temperature or surface gravity.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f20.pdf}}
\caption{
Same as Fig. \ref{co} but for neutral (left) and ionized (right)
chromium.
}\label{cr}
\end{figure}
While most of the stars in Fig. \ref{cr} for both ionization states
are below solar, the trend for [Cr/Fe] is near 0.0 dex with a number of stars
beyond 1$\sigma$ above the solar value of [Cr/Fe] = 0.0 dex. The 1$\sigma$ scatter in [Cr/Fe] is relatively constant, around 0.18 dex,
for [Fe/H] $<$ 0.2 dex, where the number of stars observed at higher metallicities severely decreases.
Comparatively, the scatter in [Cr II/Fe] is roughly half of that for the neutral state, please reference Table \ref{tab.sigma}, while the abundances for [Cr II/Fe] are markedly lower.
As discussed in \citet{Neves:2009p1804}, [Cr II/Fe] has a
slight downward trend with increasing [Fe/H], with a corresponding weak trend for [Cr/Fe]. These differences may be the result of weak, blended Cr II lines \citep{Neves:2009p1804}, different surface gravities for the two ionization states \citep{Reddy:2003p1354, Gratton:2003p1182}, or overionization from Cr II \citep{Feltzing:1998p886}.
Noticeable in both plots in Fig. \ref{cr} is an ``ensemble" of stars the show a depletion in the Cr I/II ratio and increased [Fe/H]. A significant fraction of stars that are within 60 pc of the Sun (green and yellow circles) appear to be clustered around [Cr/Fe] $\approx$ -0.2 dex and and below a gap at -0.4 dex for [Cr II/Fe]. Further investigation into this phenomena has found many of these measurements originated from \citet{Neves:2009p1804}, without which the higher [Fe/H] measurements and lower Cr I/II abundance ratio regime would be sparsely populated. Similar to a number of elements already discussed,
like silicon, sulfur, and ionized scandium, the [Cr II/Fe] vs. [Fe/H] plot shows a dual
ensemble of stars, separated by $\sim$ 0.1 dex around [Cr II/Fe] = -0.4 dex (see \S \ref{gaps} for more discussion).
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f21.pdf}}
\caption{
Same as Fig. \ref{co} but for manganese (left) and cobalt (right).
}\label{mnco}
\end{figure}
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f22.pdf}}
\caption{
Like Fig. \ref{tirad} but for manganese.
}\label{mnrad}
\end{figure}
Variation of the [Mn/Fe] ratio with [Fe/H] for the 512 stars in the analysis of
Hypatia is shown in Fig. \ref{mnco} (left). A unique element in the
iron group, the manganese and iron ratio is shown to slightly increase with [Fe/H] \citep{helfer_1959_aa,gratton_1989_aa,goswami_2000_aa,Feltzing:2007p855}.
The 1$\sigma$ scatter in [Mn/Fe] is large and variable, going from 0.17 dex at [Fe/H] = -0.4 dex, to 0.27 dex at solar [Fe/H], and then down again to 0.08 dex at [Fe/H] = 0.4 dex (see Table \ref{tab.sigma}). The majority of the scatter above the 84\% percentile regression line, particularly [Mn/Fe] $>$ 0.0 dex, is largely from \citet{Thevenin:1998p1499}, with a few contributions from other data sets. The slightly increasing median trend line, coupled with notable abundance
corrections from hyperfine splitting effects in strong lines,
has made
the rise in [Mn/Fe] with [Fe/H] challenging to decipher. Manganese could have formed from either core collapse supernovae or SN Ia, but it is difficult to tell which is dominating the trend
for [Fe/H] $>$ -1.0 dex \citep{Chen:2000p1857,Prochaska:2000p3329,
McWilliam:2003p3332,bergemann_2007_aa,
Feltzing:2007p855,bergemann_2008_aa}.
Fig. \ref{mnrad} shows [Mn/Fe] ratios versus radial distance, colored
to show the height above the Galactic plane (left) and
directionality towards or away from the galactic center (right).
There is scatter in [Mn/Fe], $\approx$ 0.7 dex, for stars that are nearer to the Sun,
although at further distances the scatter decreases to $\approx$ 0.5 dex.
Interestingly, stars near to the Sun, or within $\sim$ 30 pc, that have [Mn/Fe] $>$ -0.2 dex tend to be closer to the Galactic plane (red circles) than the majority of stars, with [Mn/Fe] $<$ -0.2 dex. There is a relative dearth of stars well above or below the plane with more enriched [Mn/Fe] ratios. Whether this is a result of survey biases or a physical causation remains to studied further.
Stars that are towards the galactic anti-center (purple circles) are more enriched in [Mn/Fe]
than the other neighboring stars, with few of them exhibiting abundances below -0.4 dex, especially at nearby radial distances.
The single stable isotope of cobalt,
$^{59}$Co, is produced by a variety of processes in several sources;
as a result, there is no consensus on the overall trend of cobalt in
halo, thick, or thin disk stars, nor a generally accepted production origin site \citep{bergemann_2010_aa}.
Variation of the [Co/Fe] ratio with [Fe/H] is shown in Fig. \ref{mnco} (right).
The median [Co/Fe] ratio indicates a relatively flat trend, that goes slightly concave-up for higher [Fe/H] abundances. This feature is made more pronounced with the inclusion of the 1$\sigma$ regression lines, and has noted by several authors
\citep{Reddy:2003p1354,del-peloso_2005_aa,Reddy:2006p1770,Neves:2009p1804}.
For [Fe/H] below solar, [Co/Fe] decreases from $\approx$ 0.0 dex down about 0.1 dex,
with a scatter varying from $\approx$ 0.1 dex to $\approx$ 0.2 dex (see Table \ref{tab.sigma}). When [Fe/H] is above solar, [Co/Fe] increases very slightly, although remaining below solar, with a scatter of $\approx$
0.15 dex. Similar to [Mn/Fe] (left), stars that are at larger distances above 60 pc are clustered
around the solar value of [Co/Fe], with few stars at greater distances measured below -0.1 dex, giving cause to further abundance surveys for stars at 60 pc or greater.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f23.pdf}}
\caption{
Similar to Fig. \ref{naalpha} but for the $\alpha$-chain elements
[Ni/Ca] (left) and [Ni/Si] (right) plotted against [Fe/H].
}\label{caalpha}
\end{figure}
Nickel has the most measurements of any element within the iron group
in Hypatia; forty literature sources are listed with [Ni/Fe]
abundances in Table \ref{tab.cat}. This is because nickel
has a similar ionization potential and atomic structure as iron and is
relatively easy to measure in the optical spectrum.
For example, \citet{Gilli:2006p2191} determined the abundances
nickel (and 11 other elements) for 101 stars in the solar
neighborhood, 93 of which are known to be host-stars to exoplanets. A
total of 98 of their stars are in the {\it Hypatia Catalog}.
Their spectra were acquired using five different
spectrographs that, in total, spanned the range of 3800 \AA\ to \hbox{10\,000
\AA}, with significant overlap in wavelength coverage between the spectrographs.
The maximum resolution was $\lambda / \Delta \lambda \approx$ 110\,000 and
minimum resolution of
$\lambda / \Delta \lambda \approx$ 48000. A standard LTE analysis,
with respect to the solar abundances determined by
\cite{Anders:1989p3165}, was done for all elements using MOOG
\citep{Sneden:1973p6104} and the ATLAS9 atmospheres
\citep{Kurucz:2005p4698}. Effective temperatures, surface gravities,
microturbulence, and metallicity [Fe/H] were all determined by
\citet{Santos:2005p2866,Santos:2004p2996}. The spectral lines that
were used for refractory elements matched those within
\citet{Bodaghee:2003p4448}, while the lines for the other elements are
from \citet{Beirao:2005p482}. \citet{Gilli:2006p2191} estimate an
overall uncertainty of of $\sim$0.10 dex for all abundance determinations.
Fig. \ref{caalpha} shows the evolution of [Ni/Ca] (left) and [Ni/Ca]
(right) with [Fe/H]. Both show positively sloped trends with [Fe/H],
indicating the injection of the iron-peak nuclei from SN Ia for all [Fe/H] abundance measurements.
The scatter of [Ni/Ca] is $\approx$ 0.2 dex, with a very small percentage of stars above or below the
16-84\% quantile regressions lines. In comparison, [Ni/Si] (Fig. \ref{caalpha}, right) has a scatter of $\sim$ 0.2 dex for [Fe/H] $\ltaprx$ 0.0 dex, which tapers to $\sim$ 0.1 dex for [Fe/H] above solar. In both plots, there is relatively small scatter about the median trend lines (blue). From Fig. \ref{fig.hist}, it is clear that all three elements are highly measured in the Hypatia Catalog. Looking at the individual abundance plots for Si (Fig. \ref{sis}), Ca (Fig. \ref{ca}), and Ni, (Fig. \ref{ni}), we see that all three have small 1$\sigma$ scatter in their own regard (Table \ref{tab.sigma}). Examining the correlations in Fig. \ref{caalpha} provides a stringent test of the Hypatia methodology. These elements are all well measured in many stars by many surveys. The spectra have numerous, strong, unblended lines in the optical, and so are less prone to systematic differences in derived abundances with different measurement techniques. Theoretically, they are expected to have a relatively tight correlation. Any errors in the methodology of combining, renormalizing, compiling, or with respect to making kinematic or other cuts to the data would be most easily detected in departures from these correlations. This gives us more confidence in the robustness of trends, such as those discussed in \S \ref{gaps}.
Evolution of the [Ni/Fe] ratio with [Fe/H] for the 1035
stars in the Hypatia analysis is shown in Fig. \ref{ni}.
The 1$\sigma$ scatter evolves from $\sim$ 0.05 dex to $\sim$ 0.20 dex to $\sim$ 0.10 dex, per Table \ref{tab.sigma}, as metallicity increases. Many stars cluster near the solar value of [Ni/Fe], as well as [Ni/Fe] = -0.2 dex. We note that small group of stars between [Ni/Fe] = [0.0, 0.2] for [Fe/H] $>$ 0.0 dex is mostly from the \citet{Trevisan:2011p6253} dataset. Similar to the other elements displaying a ``gap" in their abundance plots (Mg, Si, S, ScII, and CrII), we note that the
ensembles are spatially dependent, such that the stars further away are more enriched
in [Ni/Fe]. The lower ``ensemble" stars, at radial distances less than 60 pc, originate from a number of different datasets, including \citet{Valenti:2005p1491}. See \S \ref{gaps} for more discussion.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f24.pdf}}
\caption{
Same as Fig. \ref{co} but for nickel.
}\label{ni}
\end{figure}
\section{Beyond the Iron-Peak (Cu, Zn, Sr, Y, \& Zr)}
\label{s.beyond}
Nuclei above the iron peak have large Coulomb barriers, making
charged-particle interactions unlikely at temperatures that would not
photodissintegrate the nuclei. These elements also have
fewer spectral lines in the optical regime, resulting in fewer
abundance determinations as shown in
Fig. \ref{fig.hist}.
As a consequence, the x- and y-axis scales for the elements beyond the iron-peak show a wider range compared with the previous elements, such that [Fe/H] = [-0.7, 0.4] and [X/Fe] = [-0.5, 1.0], respectively.
Copper has two stable isotopes, $^{63}$Cu and $^{65}$Cu; the more common isotope, $^{63}$Cu, is mostly created
as radioactive $^{63}$Ni via the s-process during
hydrostatic helium burning in massive stars and intermediate mass AGB
stars.
The evolution of [Cu/Fe] with [Fe/H], as shown in Fig. \ref{cuzn} (left),
is mostly constant and sub-solar with some scatter about the blue median line. The only exception are
the two outlier stars that have [Cu/Fe] $>$ 0.4 dex; namely, HIP 78680 and HIP 64497, both measured by \citet{Ramirez:2009p1792}. Because the low number of stars for which [Cu/Fe] was measured, particularly at lower
[Fe/H] abundances, the 1$\sigma$ quantile regression lines could not be accurately determined.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f25.pdf}}
\caption{
Same as Fig. \ref{co} but for copper (left) and zinc (right).
Due to small number statistics, the percentile trend lines could not be accurately
determined for [Cu/Fe] and [Zn/Fe], which has less than 250 stellar measurements (see Table \ref{tab.skip}).
Also note the y-axis scale change.
}\label{cuzn}
\end{figure}
Zinc is made as radioactive $^{64}$Ge
in the s-process during hydrostatic helium burning
\citep{couch_1974_aa,iben_1982_aa,busso_1999_aa,kappeler_2011_aa}
and in $\alpha$-rich freezeouts from massive stars \citep{woosley_1973_aa,
hix_1996_aa},
but consensus on the origin site has not yet been reached
\citep{hoffman_1996_aa,umeda_2002_aa,Chen:2004p5708,mishenina_2011_aa}.
The evolution of [Zn/Fe] with [Fe/H], as shown in Fig. \ref{cuzn} (right), has a relatively flat trend, punctuated by a small number of measurements at
the extreme [Fe/H] metallicities.
Noted also by \citet{roederer_2010_aa, kobayashi_2011_aa}, the majority of the [Zn/Fe] abundances are sub-solar.
This is particularly interesting given the markedly non-solar abundances also seen in [Cu/Fe] with respect to [Fe/H],
shown in Fig. \ref{cuzn} (left). The sub-solar patterns for both of these elements has not been highlighted much in previous work, however, the larger number of abundances measurements in the {\it Hypatia Catalog} allows a better comparison of the trends.
Due to the difficulty of measuring
[Zn/Fe], there were not enough stars to accurately determine the 1$\sigma$ trend lines about the median,
which emphasizes the need to better understand the trends in Fig. \ref{cuzn}.
A total of 64 stars within the {\it Hypatia Catalog} have copper and
zinc abundances determined by \citet{Ramirez:2009p1792}.
Their spectra has a resolution of $\lambda / \Delta \lambda
\approx$ 60\,000 over the range 3800--9125 \AA \, and high
signal-to-noise (S/N $\approx$ 200 per spectral pixel) for the solar
twins and analog stars. They determined the solar flux spectrum by
using differential analysis from asteroid spectra. Effective
temperatures, surface gravities, and microturbulent velocities were
obtained by iterating until the difference in [Fe/H] values from both
Fe I and Fe II approached zero. Model LTE atmospheres
from \citet{Kurucz:2005p4698} and MOOG \citep{Sneden:1973p6104} were
used to determine all element abundances. The reported uncertainties
for the abundance measurements is [X/Fe] $\approx$ 0.03 dex.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f26.pdf}}
\caption{
Same as Fig. \ref{co} but for neutral (left) and ionized (right)
strontium.
Due to small number statistics, the median and/or percentile trend lines could not be accurately
determined for [Sr/Fe] and [SrII/Fe], which both have less than 250 stellar measurements (see Table \ref{tab.skip}).
}\label{sr}
\end{figure}
There are multiple processes that produce
strontium, yttrium, and zirconium. According to
\citet{Arlandini:1999p4910}, 85\% of strontium in the solar composition
is from the r-process and 15\% from the s-process.
Similarly, 92\% of yttrium and 83\% of zirconium is
through the r-process while 8\% and 17\%, respectively, is from the
s-process. While the relative contributions from each process depend
on the initial enrichment and the age of the stellar system,
there are unaccounted contributions at low metallicities
\citep{Thielemann:2007p4473}.
These abundance remainders might
be due to primary production of strontium, yttrium, and zirconium within
massive stars, possibly through neutrino-proton interactions
\citep{arnould_2007_aa}.
Nine literature sources within Hypatia measured strontium for
274 stars: \citet{Galeev:2004p979,Luck:2005p1439,Mashonkina:2007p1419,
Reddy:2003p1354,Thevenin:1998p1499}. Due to the limited number of
available lines in the optical spectrum, a number of these literature
sources quoted high uncertainties from blended, weak lines.
The abundance of [Sr/Fe] as a function of [Fe/H] is shown in Fig. \ref{sr} (left).
The median of the [Sr/Fe] abundances is plotted in blue with respect to [Fe/H].
Due to the small number of stars, the 1$\sigma$ regression lines could not be calculated,
but the scatter appears relatively large, which may be indicative
of multiple origin sites \citep{Lai:2007p4943}. The outlier stars, with [Sr/Fe] $>$ 0.4, are
a product of a number of different catalogs, such as \citet{Allen:2011p355, Thevenin:1998p1499, Luck:2005p1439}.
There are 25 stars in the analysis of the {\it Hypatia Catalog}, see \S \ref{s.struct}, which have strontium
abundances determined from Sr II, as shown in in Fig. \ref{sr} (right).
The scatter in [Sr II/Fe] is large with no clear clustering
below solar, unlike [Sr/Fe]. No trend lines could be determined for [SrII/Fe].
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f27.pdf}}
\caption{
Same as Fig. \ref{co} but for neutral (left) and ionized (right)
yttrium.
Due to small number statistics, the percentile trend lines could not be accurately
determined for [Y/Fe] and [YII/Fe], which have less than 250 stellar measurements (see Table \ref{tab.skip}).
}\label{y}
\end{figure}
Many catalogs within Hypatia made the distinction between measuring
neutral or singly ionized yttrium, although only \citet{Feltzing:1998p886}
measured both Y I and Y II. Evolution of [Y/Fe] as a function of
[Fe/H] for 191 stars is shown in Fig. \ref{y} (left)
while [Y II /Fe] versus [Fe/H] for 210 stars is shown on the right.
The more numerous Y II measurements are due to
rather weak and blended lines available for Y I in the optical
spectrum.
The median [Y/Fe] abundance trend, represented
by the blue line, suggests a flat and slightly sub-solar trend
\citep{qian_2008_aa,roederer_2010_aa},
but with a relatively large dispersion for most of [Fe/H]. The stars with super-solar [Y/Fe], or greater than 0.5 dex, were measured by a number of catalogs, such as \citet{Thevenin:1998p1499, Wang11, Allen:2011p355}.
\citet{Wang11} used a mixture of both Y I and Y II lines in their abundance determinations, while the abundances in \citet{Allen:2011p355} employed differing measurement techniques between lines (see their Table 2). The line list used by \citet{Thevenin:1998p1499} was not published. Due to the abnormality of these super-solar [Y/Fe] abundances, these results may be due to varying reduction methods.
Until further investigation, the validity of their results remains tentative.
There are relatively few stars with
a radial distance less than 30 pc
and the majority of those stars exhibit near-solar or below [Y/Fe] abundances.
Comparatively, the evolution the [Y II/Fe] ratio with [Fe/H] is
also relatively flat and below-solar. However, the dispersion is much less than [Y/Fe], most likely due to the smaller number of unblended Y II lines. This suggests that the smaller scatter in the [YII/Fe] measurements are more likely to be physical.
Zirconium abundances in Hypatia are similar to those of yttrium, in
that the singly ionized state was preferentially measured due to
blending of weak neutral lines in the optical spectrum.
Evolution of [Zr/Fe] as a function of [Fe/H] for the 96 stars in the analysis of Hypatia (see \S \ref{s.struct}) is shown
in Fig. \ref{zr} (left), and [Zr II /Fe] versus [Fe/H] for 188
stars is given in Fig. \ref{zr} (right).
Because both Zr and Zr II have been measured than less than 250 stars, their 1$\sigma$ quantile regression trends couldn't be determined. The abundances for [Zr/Fe] are generally solar \citep{qian_2008_aa,kashiv_2010_aa}, although with significant scatter that makes the trend difficult to discern.
There are 14 stars with [Zr/Fe] $>$ 0.4 dex, contributed from \citet{Thevenin:1998p1499, Allen:2011p355}. As discussed above, the varying techniques employed by within these datasets may have contributed to the widely varying [Zr/Fe] measurements.
There are few abundance measurements around [Fe/H] $<$ -0.4 dex, as noticed by the sharp peak in the median line.
Comparatively [Zr II/Fe] shows a shallow concave-up trend and less scatter, since the two plots in Fig. \ref{zr} have the same axes. The outlier stars, with [ZrII/Fe] $>$ 0.25 originate from both
\citet{Reddy:2003p1354,Allen:2011p355}.
Similarities between the patterns seen in [Zr II/Fe] and [Y II/Fe]
with respect to [Fe/H] suggest similar origin sites.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f28.pdf}}
\caption{
Same as Fig. \ref{co} but for neutral (left) and ionized (right)
zirconium.
Due to small number statistics, the percentile trend lines could not be accurately
determined for [Zr/Fe] and [ZrII/Fe], which have less than 250 stellar measurements (see Table \ref{tab.skip}).
}\label{zr}
\end{figure}
\citet{Bond:2006p2098} determined Y II and Zr II abundances in of 144
G-type stars, all of which are in Hypatia, using high resolution
($\lambda / \Delta \lambda \approx$ 80\,000), high signal-to-noise ratio
(S/N $\approx$ 250 per pixel) spectra of the 5380.32 \AA \ and 6587.62
\AA \ lines of C I. Oscillator strengths and excitation energies for
each line were obtained from the NIST Atomic Spectra
Database. Effective temperatures were determined from the stellar
colors listed in the {\it Hipparcos Catalogue}, which is different
from similar studies that utilize the spectra for an effective
temperature value, and resulted in a mean $T_{eff}$ uncertainty of
$\pm$ 100 K. An LTE analysis using WIDTH6
\citep{Kurucz:2005p4698}
in conjunction with a grid of \citet{Kurucz:2005p4698} ATLAS9 atmospheres was
used to determine the elemental abundances. Surface gravities $\log
g$ values and Fe abundances were obtained by iterating until the
[Fe/H] value from both Fe I and Fe II was the same. The
microturbulence parameter $\zeta$ which minimized the correlation
coefficient between $\log \zeta$ for Fe I and $\log (W_{\lambda} /
\lambda$) was selected with an estimated uncertainty of $\pm$ 0.25.
\section{Neutron-Capture Elements (Ru, Ba, La, Ce, Pr,\\ Nd, Sm, Eu, Gd, Dy, \& Pb)}
\label{s.neutron}
The heavy neutron capture elements have the majority of their lines in the
extreme blue and UV part of the spectrum, making it challenging to
measure these elements \citep{spite_1978_aa,sneden_2008_aa}. As a
result, fewer catalogs report these abundances (see
Fig. \ref{fig.hist}) and the observational uncertainties in the
abundance ratios are generally larger for the neutron capture elements.
Because of the larger scatter seen in the data, the plots shown for the neutron-capture elements have x- and y-axis scales where that [Fe/H] = [-0.7, 0.4] and [X/Fe] = [-0.6, 1.0], respectively.
We elect not to discuss Ru, La, Pr, Sm, Gd, Dy and Pb in
this paper, although these elements are listed in Hypatia, focusing
on those elements with more measurements: Ba II, Ce, Nd, and Eu.
Elemental barium is dominated by three isotopes: $^{135}$Ba, $^{137}$Ba, and $^{138}$Ba, the majority of which are made in the s-process \citep{Arlandini:1999p4910, Travaglio:1999p4899, Mashonkina:2000p4947, carlson_2007_aa}. Only singly ionized barium lines are available in the optical band for FGK-type stars, which are strongly affected by hyperfine splitting \citep{Mashonkina:2000p4947, Mashonkina:2007p1419}. Variation of [Ba II/Fe] with [Fe/H] for the 301 stars in the analysis of Hypatia is shown in Fig. \ref{bace} (left). The median and 1$\sigma$ regression lines suggest a relatively flat and slightly sub-solar [Ba II/Fe] ratio with a large scatter of $\approx$ 0.26 dex (see Table \ref{tab.sigma}). There are also a number of outlier stars with extremely enriched Ba II at distances greater than 30 pc, some of which were determined by \citet{Allen:2011p355}. Similar to yttrium and zirconium, the stars that are closer to the Sun (green circles) have higher [Fe/H] content as well as near- or below-solar values of [Ba II/Fe].
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f29.pdf}}
\caption{
Same as Fig. \ref{co} but for ionized barium (left) and cerium (right).
Due to small number statistics, the percentile trend lines could not be accurately
determined for [Ce/Fe], which has less than 250 stellar measurements (see Table \ref{tab.skip}).
}\label{bace}
\end{figure}
Cerium is also predominantly (77\%) made by made by the s-process as
either $^{140}$Ce or $^{142}$Ce \citet{Arlandini:1999p4910,simmerer_2004_aa, kappeler_2011_aa}.
Twelve catalogs within Hypatia measured cerium in 237 stars, due to only
one or two lines in the optical spectrum that were not strongly
affected by blending \citep{Luck:2005p1439}.
Fig. \ref{bace} (right) shows a predominantly near-solar [Ce/Fe] trend with respect to
[Fe/H], as highlighted by the median curve, also noted by \citet{Brewer:2006p1310}. Given the small number of stars, the scatter was difficult to determine using the 1$\sigma$ quantile regression analysis. However, similar to [BaII/H], \citet{Allen:2011p355}, as well as \citet{Thevenin:1998p1499}, reported multiple extremely high [Ce/Fe] abundance ratios for the more distant stars.
Determinations for a wide-variety of neutron-capture elements were conducted by
\cite{Galeev:2004p979}, who
measured Ru, Ba II, La, Ce, Pr, Nd, Sm, and Eu
abundances for 15 stars in the solar neighborhood, all of which are in
the {\it Hypatia Catalog}, using the 2 m reflector at
Terskol Observatory in the northern Caucasus. Their high resolution
spectra ($\lambda / \Delta \lambda \approx$ 45000) covered 4000 - 9000\AA
with a signal-to-noise ratio of 150-200 per spectral pixel. They
determined the solar flux spectrum via differential analysis,
using solar light scattered off of the Earth's atmosphere. An LTE
analysis with WIDTH6 \citep{Kurucz:2005p4698}
is used, in conjunction with oscillator strengths
from the VALD database \citep{Kupka:1999p4972}, to determine the final
elemental abundances. An accuracy of 0.10 dex is assigned to their [Fe/H] measurements.
The seven stable isotopes of neodymium are made by both the r- and
the s-processes \citep{Roederer:2008p4949}.
Due to blending with nearby
lines, there are only one or two lines in the optical spectrum from
which [Nd/Fe] is measured \citep{Galeev:2004p979,Bond:2008p2099}.
The median [Nd/Fe] abundance, shown by the blue line,
indicates a relatively flat, solar trend for [Nd/Fe] with [Fe/H], see Table \ref{tab.sigma} as well as
\citet{Thevenin:1998p1499,Reddy:2003p1354}. Due to the small number of stars for which
this element was measured, as well as the large scatter (which was not due to directly to any one
catalog), it is difficult to discern any robust trend.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f30.pdf}}
\caption{
Same as Fig. \ref{co} but for neodymium (left) and europium (right).
Due to small number statistics, the percentile trend lines could not be accurately
determined for [Nd/Fe] and [Eu/Fe], which have less than 250 stellar measurements (see Table \ref{tab.skip}).
}\label{ndeu}
\end{figure}
About 91\% of europium is estimated to be from the r-process
\citep{Cameron:1982p4965,Arlandini:1999p4910} and is often used as a
standard against a predominantly s-process element, such as [Ba/Eu]
\citep{mcwilliam_1997_aa, Travaglio:1999p4899, Mashonkina:2000p4947}.
There are eighteen literature sources within Hypatia that measured
[Eu/Fe] using one or two lines in the optical spectrum.
Unlike neodymium, the trend of [Eu/Fe] with respect to [Fe/H] has a negative
slope with [Fe/H] \citep{Thevenin:1998p1499,Reddy:2003p1354,
Galeev:2004p979,Bond:2008p2099,Allen:2011p355}, as shown in
Fig. \ref{ndeu} (right).
However, the overall scatter away from the median trend line is relatively large,
$\approx$ 0.8 dex over the range of [Fe/H]
shown. There were few [Eu/Fe] measurements for stars at distances near to the Sun. In general, the abundance measurements (most of which were from stars greater than 30 pc) tend to be clustered at [Eu/Fe] $\approx$ 0.0 dex and [Fe/H] $\approx$ -0.2 dex \citep{Thevenin:1998p1499,Allen:2011p355}.
\section{Gaps in Abundances}
\label{gaps}
We found a number of elements, specifically Mg, Si, S, Sc II, Cr II, and Ni, that displayed systematic correlations, namely two ensembles separated by a ``gap" in their abundance ratios relative to Fe. In each case, the ensemble with a higher [X/Fe] tends to be at lower [Fe/H], though there is overlap. In all cases, the ensembles showed a bias with respect to stellar distance. Stars below the gap are closer, on average than the stars above. In some cases the majority of the stars contributing to the lower ensemble were attributable to a single survey. However, the dominant survey was different in the cases of different elements (i.e. \citet{Neves:2009p1804} for Mg, \citet{Valenti:2005p1491} for Si). Furthermore, when the survey was removed, the distinct ensembles were still identifiable, but with smaller statistics. These were often stars with abundances measured in multiple surveys. Further investigation revealed that the surveys that contributed large numbers of stars to the lower ensemble, for example \citet{Neves:2009p1804, Valenti:2005p1491}, analyzed more nearby stars. That being said, the accumulation of a number of datasets that reveal this trend, including those listed in each respective subsection, for example \citet{Sadakane:2002p6580,Neves:2009p1804}, leads us to believe that what we are seeing is two physically distinct groups of stars. We did not include Ca and Ti in this list as they do not have the immediately obvious gap shown by the elements listed above. We do, however, examine their statistics, since they are products of $\alpha$-chain burning like Mg, Si, S, and Cr. Ca does show a less prominent gap. There is only a bare suggestion of a gap in Ti, but the stair-step morphology of the plot at 0.0 $<$ [Fe/H] $<$ 0.2 is similar to the other elements considered. In both cases the division is close to [X/Fe] = -0.1, as with the other elements.
One simple demonstration of the reality of the ``gaps" is the direct analysis of two stellar spectra, with roughly equally [Fe/H] but differing elemental abundances.
In Figure \ref{spectra}, we show spectra from HD 109591(blue) and HD 50255 (pink). To the left and middle, we present two Fe I lines, where the percent difference between the equivalent width determinations per \citet{PaganoPhD} is 0.069\% and 1.6\%, respectively. For the three Ni I lines on the right, the difference between the equivalent widths are 7.5\%, 10.4\%, and 11.8\%, respectively. \citet{PaganoPhD} also derived the stellar parameters and abundances for these stars:
$T_{eff}$ = 5784, 5788 K; [Fe/H] = -0.02, -0.04 dex; and [Ni/Fe] = 0.01, -0.14 dex, respectively. Notably, while the effective temperatures and overall metallicity are the same, the nickel content (both in the raw spectra and reduced abundance ratio) vary dramatically between the two stars.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.3in]{f31.pdf}}
\caption{
Continuum normalized spectra for two stars with nearly equal iron abundance, see lines on the left and middle, but noticeably differing nickel abundances, see three lines on the right \citep{PaganoPhD}.
}\label{spectra}
\end{figure}
To better understand the stars within each ensemble above and below the gap, we analyzed the z-height distribution with respect to the Galactic plane. In Fig. \ref{zht}, we show the stars with silicon measurements (left) and nickel abundances (right), with exponential fits to the distributions. There is overlap between the ensembles, with both having members at all heights above the disk, but it is clear that the ensemble below the gap is more concentrated towards the mid-plane of the disk. Alternatively, we also examine the height $h_{90}$ below which 90\% of the stars in each ensemble can be found. Table~\ref{scale.tab} gives the scale heights of the exponential fits $h_a$ and $h_b$ for the ensembles above and below the gap, respectively, along with $h_{90a}$ and $h_{90b}$. For the definitive gap elements, in all cases $h_a$ falls between 35 and 40pc. Excluding sulfur, $h_b$ shows a slightly larger range, from 21 to 28pc. The scale height for sulfur below the gap is 34pc, but there are fewer than 60 stars upon which to base the fit. The 90\% limits show a similar pattern with a moderately larger range. Values for $h_{90a}$ vary from 68 to 80pc. Again excepting sulfur, $h_{90b}$ ranges from 36 to 50pc. The below the gap ensemble for S is closer to but still lower than the ensemble above the gap. The cases of Si and Ni, shown in Figure~\ref{zht}, actually show the least difference between the vertical distributions of the ensembles if S is removed. In both cases $h_b$ is 70\% of $h_a$, and $h_{90b}$ 61\% and 63\% of $h_{90a}$, respectively. The addition of the potential gap elements changes the picture very little. Calcium is solidly within the normal distribution for both $h$ and $h_{90}$. The separation in Ti is relatively small, with $h_a$ and $h_b$ differing by 3pc, though there is still a difference of 21pc in $h_{90}$.
\begin{figure}[ht]
\centering
\centerline{\includegraphics[height=2.5in]{f32.pdf}}
\caption{
Sorted z-height data above and below the respective gaps for stars with silicon (left) and nickel (right) abundances. The legend gives the scale height (H) as a result of the fitted exponentials.
}\label{zht}
\end{figure}
For an additional test regarding the contribution of individual datasets to Hypatia, we recalculated the scale-heights without the inclusion of the \citet{Neves:2009p1804} dataset, one of the largest contributors for many of the lower ``ensembles." We then recalculated the scale-heights to better understand the impact of their survey, see Table \ref{scale2.tab}.
We found that the scale-heights did not vary more than 2pc for either ensemble in Mg, Si, S, and Ni, or for the ensemble above the gap in Ca and Cr. The $h_a$ for Sc dropped by 5pc, with roughly half the number of stars as the full sample. The $h_a$ for Ti increased by 4pc, actually making the ensembles above and below more distinct. The value of $h_b$ changed by less than 2pc for all elements except Ca and Cr. In both cases there were fewer than 20 stars remaining below the gap when the survey was removed. The situation was similar for $h_{90}$. In every case but two (6pc for Sc and 4pc for Ti) $h_{90a}$ changed by 2pc or less. The value of $h_{90b}$ changed by no mare than 2 pc for all of the elements except Ca, Sc, and Cr. In all of these cases the number of stars remaining below the gap was $\ltaprx$ 20. We would expect this measure to be very sensitive to small number statistics. In general, we found that the structure of the two ``ensembles" is unaffected by an individual catalog.
This set of tests shows remarkable consistency, and suggests that the stars above and below the gap may in fact be physically distinct ensembles. We find that there may be an inhomogeneity in the solar neighborhood stars, specifically a group near to the mid-plane that is high in [Fe/H] but deficient in the elements Mg, Si, S, Ca, Sc II, Ti, Cr II, and Ni. The case is particularly strong for Mg, Si, S, Ca, Sc, Cr, and Ni. Titanium shows a similar overall morphology, though the gap is at least partially filled. The others, Mg, Si, S, and Ca, are $\alpha$-elements, and the dominant isotopes of Ti and Cr are direct products of $\alpha$-chain burning ($^{48}$Ti and $^{52}$Cr are beta decay products of $^{48}$Cr and $^{52}$Fe, respectively.) Nickel is a primarily product of $\alpha$-captures onto $^{54}$Fe and $^{56}$Fe in a moderately neutron rich environment. Scandium is an odd-z element produced mostly in core-collapse SNe. It would be instructive to look for similar behavior in the nearby odd-z elements V and Mn, but these have extremely high spreads, perhaps due to incorrect handling of hyperfine corrections as noted above. Contributions from SNIa may be partly obscuring the gap in Ti in particular, and to an extent in Ca. We note that the gap occurs most clearly for elements that are dominated by a single nucleosynthetic production site, explosive O/Si burning or $\alpha$-rich freezeout in core-collapse SNe. Other nucleosynthetic sources or observational vagaries may be obscuring a gap that would otherwise appear for other elements. It will prove valuable to explore this possible nearby asymmetric abundance distribution, particularly as it indicates that the solar neighborhood is not well-mixed compositionally. Information on ages and detailed kinematics of these ensembles could prove highly interesting.
\section{Discussion}
\label{disc}
We have assembled spectroscopic abundance data from 84 literature
sources for 50 elements across 3058 stars within 150 pc of the Sun to
build the {\it Hypatia Catalog}. This is the largest, most complete
catalog of spectroscopic abundance data for stars in the solar
neighborhood to date, of which we are aware (see Fig.\,\ref{fig.hist}). The aim of this research is to (a) provide an unbiased compilation of abundance determinations
of main-sequence stars in the solar neighborhood; (b) show the zeroth order abundance
trends within Hypatia are consistent with the published literature; and (c) explicitly show
the challenges of combining disparate data sets in hopes for some reconciliation within the community.
We encountered a number of issues by trying to amalgamate such
large and diverse data sets, which we attempted to ameliorate in the most unbiased manner.
We began by first determining the likely origin for the stars within the {\it Hypatia Catalog} using the prescription in \citet{Bensby:2003p513}, in order to exclude thick-disk stars from our analysis.
We then undertook to minimize the spread, or the variation in abundance
measurements for the same star and element reported by different literature sources (see
Fig.\,\ref{spread}), and renormalizing all the stars in Hypatia to the
\cite{Lodders:2009p3091} solar abundance scale. We found that while this has
a significant impact to the abundances (0.6 dex on average), it did not significantly
affect the spread in the data.
Finally, we only retained those stars in our analysis whose spread in both [X/Fe] and [Fe/H] was
less than the respective error bar. In this way, we were able to ensure that
the abundances we studied were agreed upon by multiple datasets, were well understood, and were
without controversy (\S \ref{s.struct}).
We found abundance trends that are consistent with previously
discovered average trends. In addition, the large number of stars in Hypatia
allows us to observationally quantify the extent of the scatter, or
the width about the median abundance trend using quantile regression for the 16\% and 84\% (total 68\% or 1 standard deviation) percentiles,
for each element
\citep[Figs.\,\ref{co}-\ref{ndeu}, and also see][]{tinsley_1980_aa,malinie_1993_aa,
van-den-hoek_1997_ab,kobayashi_2011_aa}.
As a result of our careful analysis, we find that the solar neighborhood may not be homogeneously well-mixed. A number of elements revealed two distinct ``ensembles" of stars with a possible asymmetric abundance distribution. Stars that were consistently near to the Galatic mid-plane were deficient in Mg, Si, S, Ca, Sc II, Cr II, and Ni but enriched in Fe, as opposed to stars further from the mid-plane, which implies a possible variation in the nearby chemical history.
One of the problems, when analyzing the abundance trends in the majority of literature sources,
is that few data sets have enough stars to discern obvious patterns that are statistically viable.
The {\it Hypatia Catalog} contains a large amount of both kinematic and
abundance data. Some applications of this data include
stellar abundance trends between thin and thick disk stars, slow and fast rotating
stars, stars of different spectral types (or effective temperatures),
exoplanet hosts versus stars confirmed to be
without exoplanets, or solar
analog stars. Hypatia may also be used to supplement pre-existing
surveys such as NASA's Transiting Exoplanet Survey Satellite (TESS) mission, the ESA/NASA Herschel mission, or
the Sloan Digital Sky Survey. None of the methods that we undertook in our analysis
are reflected in the published version of the {\it Hypatia Catalog}, which contains
only the data as determined by the original authors. In this way, others may analyze whatever datasets they choose in the way they see fit, making Hypatia an excellent resource. The breadth of information present within the Hypatia
Catalog also makes it useful for gleaning new stellar information, such as stellar
ages and the kinematics of the solar neighborhood.
\section{Acknowledgments}
\acknowledgments
The authors thank John Shumway, Scott Ransom, Sandra Schmidt, Matt
Mechtley, Mark Richardson, Themis Athanassiadou, Jade
Bond, Chris Sneden, and Caleb Wheeler for helpful discussions and
computer support. The authors are grateful for the
VIZIER, ADS, and SIMBAD databases. This
work is supported by the NASA Astrobiology Institute under Grant
08-NAI5-0018 "Follow the Elements," and by the NSF under Grant PHY
02-16783 for the Frontier Center ``Joint Institute for Nuclear
Astrophysics''. Turnbull thanks Ariel Anbar for the opportunity
to join the NASA NAI ``Follow the Elements'' team.
\clearpage
\begin{table}[h]\scriptsize
\caption{Example of the Full Hypatia Catalog}
\begin{center}
\begin{tabular}{l}
Star: HIP = 400\\
HD = 225261\\
BD = B+22 4950\\
Spec Type = G9V\\
dist (pc) = 26.39 \\
RA/Dec = (1.23, 23.27)\\
Disk component: thin \\
NaH -0.31 [Valenti \& Fischer (2005)]\\
SiH -0.23 [Valenti \& Fischer (2005)]\\
TiH -0.22 [Valenti \& Fischer (2005)]\\
FeH -0.23 [Valenti \& Fischer (2005)]\\
NiH -0.4 [Valenti \& Fischer (2005)]\\
OH 0.02 [Petigura \& Marcy (2011)]\\
FeH -0.23 [Petigura \& Marcy (2011)]\\
\end{tabular}
\label{tab.hyp}
\end{center}
\end{table}
\clearpage
\begin{center}\scriptsize
\begin{longtable}{| p{3.9cm} | p{0.7cm} | p{10.0cm} |}
\hline
\bf{Literature Reference} & \bf{Stars} & \bf{Elements} \\
\hline
\hline
\citet{Allen:2011p355} & 33 & (Fe, Mn, Cu, Zn, Y, Ba II, Nd, Eu, Gd, Dy) \\
\citet{AllendePrieto:2004p476} & 118 & (C, O, Mg, Si, Ca II, Sc II, Ti II, Fe, Co, Ni, Cu, Zn, Y, Ba II, Ce, Nd, Eu) \\
\cite{Bensby:2005p526} & 144 & (Fe, Na, Mg, Al, Si, Ca, Ti, Cr, Ni, Zn, Y II, Ba II) \\
\citet{Bergemann10} & 8 & (Fe, Cr, Cr II, Mg) \\
\citet{Bodaghee03} & 120 & (Fe, Si, Ca, Sc II, Ti, V, Cr, Mn, Co, Ni) \\
\citet{Boesgaard11} & 52 & (Fe, O, Be, Ti, Mg) \\
\cite{Bond:2006p2098, Bond:2008p2099} & 144 & (Fe, C, Na, Al, Si, Ca, Ti, Ti II, Ni, O, Mg, Cr, Ba II, Y II, Zr II, Eu, Nd) \\
\citet{Brewer:2006p1310} & 531 & (O, Na, Mg, Al, Si, Ca, Ti, Sc, V, Cr, Mn, Fe, Co, Ni, Cu, Zn, Sr, Y, Zr, Ba II, La, Ce, Nd, Eu) \\
\citet{Brugamyer:2011p3104} & 121 & (Fe, Si, O) \\
\citet{Caffau:2011p3195} & 20 & (Fe, P) \\
\citet{Carretta:2000p6235} & 19 & (Fe, C, N, O, Na, Mg) \\
\citet{Castro97} & 4 & (Na, Mg, Si, Ca, Ti, V, Ni, Zr, Y, Ba II, Eu II, Fe) \\
\citet{Castro:1999p230} & 13 & (Fe, Cu, Ba II) \\
\citet{Chen:2000p1857} & 88 & (Fe, O, Na, Mg, Al, Si, K, Ca, Ti, V) \\
\citet{daSilva12} & 25 & (Fe, C, Na, Mg, Si, Ca, Sc, Ti, V, Cr, Mn, Co, Ni, Cu, Zn, Sr, Y II, Zr, Ba II, Ce II, Nd II, Sm II) \\
\citet{DelgadoMena10} & 369 & (Fe, O, Ni, C, Mg, Si)\\
\citet{DOrazi12} & 6 & (Fe, Y II, Zr II, Ba II, La II, Ce II) \\
\citet{Ecuvillon:2004p2198} & 126 & (Fe, Zn, Cu, C, S) \\
\citet{Ecuvillon:2006p8109} & 93 & (Fe, O) \\
\citet{Edvardsson:1993p2124} & 180 & (Fe, O, Na, Mg, Al, Si, Ca, Ti, Ni, Y II, Zr II, Ba II, Nd) \\
\citet{Feltzing:1998p886} & 45 & (O, Na, Mg, Al, Si, Ca, Sc, Sc II, Ti, V, V II, Cr, Cr II, Mn, Fe, Co, Ni, Y, Y II, Zr, Mo, La, Nd, Eu, Hf) \\
\citet{Feltzing:2007p855} & 95 & (Fe, Mn)\\
\citet{Francois:1986p2312} & 36 & (Fe, Al, Si, Na, Mg) \\
\citet{Francois:1988p2318} & 11 & (Fe, S) \\
\citet{Fulbright:2000p2188} & 166 & (Fe, Na, Mg, Al, Si, Ca, Ti, V, Cr, Ni, Y II, Zr II, Ba II, Eu) \\
\citet{Galeev:2004p979} & 15 & (Li, C, N, O, Na, Mg, Al, Si, S, K, Ca, Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn, Sr, Y, Zr, Mo, Ru, Ba II, La, Ce, Pr, Nd, Sm, Eu) \\
\citet{Gebran:2010p6243} & 28 & (C, O, Na, Mg, Si, Ca II, Sc II, Fe, Ni, Y II) \\
\citet{Gilli:2006p2191} & 98 & (Si, Ca, Sc II, Ti, V, Cr, Mn, Co, Ni, Fe) \\
\citet{Gonzalez:2001p3080} & 22 & (Fe, C, O, Na, Al, Si, Ca, Sc, Ti, Ni) \\
\cite{Gonzalez:2007p1060} & 31 & (Fe, Li, C, N, Al, Ca, Mg, Na, S, Sc, Si, Ti, Cr, Cu, Mn, Ni, Zn, Eu) \\
\citet{GonzalezHernandez:2010p7714} & 95 & (Fe, C, O, S, Na, Mg, Al, Si, Ca, Sc, Ti, V, Cr, Mn, Co, Ni, Cu, Zn, Sr, Y, Zr, Ba II, Ce, Nd, Eu) \\
\citet{Gratton:2000p1209} & 58 & (Fe, C, N, O, Na) \\
\citet{Gratton:2003p1182} & 116 & (Fe, O, Na, Mg, Si, Ca, Ti, Ti II, Sc II, V, Cr, Cr II, Mn, Ni, Zn) \\
\citet{Gustafsson:1999p8407} & 80 & (Fe, C, O) \\
\citet{Huang05} & 22 & (C, O, Na, Mg, Al, Si, S, Ca, Sc II, Ti, V, Cr, Mn, Ni, Ba II, Fe) \\
\citet{Jonsell:2005p1298} & 43 & (Fe, O, Na, Mg, Al, Si, Ca, Sc II, Ti, V, Cr, Ni, Ba II)\\
\citet{Kang11} & 51 & (Fe, Na, Mg, Al, Si, Ca, Sc, Ti, T III, V, Cr, Mn, Co, Ni) \\
\citet{Koch:2002p1283} & 74 & (Eu, Fe) \\
\citet{Korotin11} & 172 & (Fe, Ba II) \\
\citet{Laird:1985p1923} & 116 & (Fe, C, N) \\
\citet{Luck:2005p1439} & 110 & (C, O, Na, Mg, Al, Si, S, Ca, Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, \break Zn, Sr, Y, Zr, Ba II, La, Ce, Pr, Nd, Sm, Eu) \\
\citet{Mashonkina11} & 25 & (Fe, Ba II) \\
\citet{Mashonkina:2007p1419} & 67 & (Fe, Sr, Y II, Zr II, Ba II, Ce) \\
\citet{Mishenina:2003p1368} & 95 & (Fe, O, Na) \\
\citet{Mishenina:2004p1360} & 173 & (Fe, Mg, Si, Ni) \\
\citet{Mishenina:2008p1380} & 129 & (Fe, O, Mg, Si, Ti) \\
\citet{Mishenina11} & 142 & (Fe, Na, Al, Cu, Zn) \\
\citet{Neuforge97} & 2 & (C, N, O, Al, Si, Ca, Sc II, Ti, V, Cr, Cr II, Mn, Fe, Co, Ni, Y II, Zr II, Eu II) \\
\citet{Neves:2009p1804} & 443 & (Fe, Si, Ca, Sc, Sc II, Ti, Ti II, V, Cr, Cr II, Mn, Co, Ni, Na, Mg, Al) \\
\citet{Nissen:1997p6199} & 19 & (Fe, O, Mg, Si, Ca, Ti, Cr, Ni, Na, Y, Ba II) \\
\citet{Nissen:2010p1903} & 43 & (Fe, Na, Mg, Si, Ca, Ti, Cr, Ni) \\
\citet{Nissen11} & 36 & (Fe, Mn, Cu, Zn, Y II, Ba II) \\
\citet{Petigura:2011p3263} & 914 & (Fe, C, O) \\
\citet{PortodeMello08} & 2 & (Fe, Na, Mg, Si, Ca, Sc, Ti, V, Cr, Mn, Co, Ni, Cu, Y II, Ba II) \\
\citet{Ramirez:2007p1819} & 523 & (Fe, O) \\
\citet{Ramirez:2009p1792} & 64 & (Fe, C, O, Na, Al, Si, S, Ca, Sc, Ti, V, Cr, Mn, Ni, Cu, Zn, Y II, Zr II, Ba II) \\
\citet{RecioBlanco12} & 9 & (Fe, Y, Zr) \\
\citet{Reddy:2003p1354} & 179 & (Fe, C, N, O, Na, Mg, Al, Si, S, K, Ca, Sc II, Ti, V, Cr II, Mn, Co, Ni, Cu, Zn, Sr, Y II, Ba II, Zr II, Ce, Nd, Eu) \\
\citet{Reddy:2006p1770} & 171 & (Fe, C, O, Na, Mg, Al, Si, Ca, Sc II, Ti, V, Cr II, Mn, Co, Ni, Cu, Zn, Y II, Ba II, Ce, Nd, Eu) \\
\citet{Sadakane02} & 12 & (C, N, O, Na, Mg, Al, Si, S, K, Ca, Sc, Sc II, Ti, T III, V, Cr, Cr II, Mn, Fe, Co, Ni, Cu, Zn, Sr, Y II, Ba II, Ce II, Nd II, Eu II) \\
\citet{Schuler11} & 10 & (C, N, O, Na, Mg, Al, Si, S, Ca, Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Zn) \\
\citet{Shi:2004p2295} & 97 & (Fe, Na) \\
\citet{Takeda:2005p6195} \, \, \, \, \, \, \, \, \, \, \, \, \, and \citet{Takeda:2007p1531} & 159 & (Fe, C, N, O, Na, Mg, Al, Si, S, Ca, Sc, Sc II, Ti, Ti II, V, V II, Cr, Cr II, Mn, Co, Ni, Cu, Zn) \\
\citet{Thevenin:1998p1499} & 663 & (Li, O, Na, Mg, Al, Si, Ca, Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Sr, Y, Zr, Mo, Ba II, La, Ce, Nd, Sm, Eu) \\
\citet{Trevisan:2011p6253} & 64 & (Fe, Ca, Si, Ti, C, Ni, O, Mg) \\
\citet{Valenti:2005p1491} & 1002 & (Na, Si, Ti, Fe, Ni) \\
\citet{Wang11} & 37 & (Fe, Al, Mg, Si, K, Ca, V, Cr, Sc II, Y, Ba II, Eu II, La II, Ni, Ti) \\
\citet{Zhang:2006p6230} & 31 & (Fe, O, Mg, Si, Ca, Ti, Na, Al, Sc, V, Cr, Mn, Ni, Ba II)\\
\citet{Zhao02} & 12 & (Fe, C, O, Na, Mg, Al, Si, S, K, Ca, Sc II, Ti, V, Cr, Mn, Ni, Ba II) \\
\hline
\caption{Literature sources used in Hypatia, with the number of stars that matched the criteria and element abundances measured.}\label{tab.cat}
\end{longtable}
\end{center}
\begin{center}\tiny \begin{landscape}
\begin{longtable} {|p{2.0cm}|p{4.0cm}|p{1.0cm}|p{1.0cm}|p{1.4cm}|p{1.5cm}|p{1.7cm}|p{2.0cm}|p{2.0cm}|p{1.0cm}|}
\hline
\bf{Catalog} & \bf{Telescope} & \bf{Resolu- tion ($\Delta \lambda /
\lambda$)} & \bf{Signal-to-Noise} & \bf{Wavelength Range (\AA)} & \bf{Stellar Atmosphere } & \bf{Equivalent Width} & \bf{Curve of Growth or Spectral Fitting} & \bf{Solar Abundance } & \bf{Number of FeI/II lines} \\
\hline
\hline
\citet{Allen:2011p355} & 1.52m telescope at European Southern Observatory (ESO), using the Fiber Fed Extended Range Optical Spectrograph (FEROS) & 45000 & 100-440 & 3560-9200 & MARCS per \citep{Gustafsson:1975p4658} & IRAF splot & ABON via \citet{Spite1967} (and improvements in the last 30yrs) & \citet{Anders:1989p3165} & 292/38 \\
\citet{AllendePrieto:2004p476} & 2.7m telescope at the McDonald Observatory using the 2dcoude spectrograph, 1.52m telescope at ESO using FEROS & 50\,000 / 45000 & 150-600 & 3600-5100 / 3500-9200 & MARCS & spectral fitting & spectral fitting & differential analysis & N/A \\
\cite{Bensby:2005p526} & 1.52m telescope at ESO using FEROS & 48000 & $>$ 300 & 3560-9200 & MARCS & IRAF splot & Uppsala EQWIDTH & their own & 176/36 \\
\citet{Bodaghee03} & 1.2m telescope at ESO using CORALIE & 50\,000 & 150-350 & 3800-6800 & ATLAS9 per \citet{Kurucz1993} & IRAF splot & MOOG per \citet{Sneden:1973p6104} & \citet{Anders:1989p3165} & 40 / 7 \\
\citet{Boesgaard11} & 10m telescope at Keck using the High Resolution Echelle Spectrometer (HIRES) & 42000 & Median 106 & 3000-6000 & ATLAS9 & IRAF splot & MOOG & their own & 48/5 \\
\cite{Bond:2006p2098, Bond:2008p2099} & 3.9m telescope at Anglo-Australian Telescope (AAT) using the University College London Echelle Spectrograph (UCLES) & 80\,000 & 200-300 & 4820-8420 & ATLAS9 & IRAF splot & MOOG & $\log$ $\sigma$(Fe) = 7.49 else \citet{Grevesse:1998p3102} & 39/6\\
\citet{Brewer:2006p1310} & 4m telescope at Kitt Peak using the echelle spectrograph & 32000 & 100-280 & 3900-7400 & ATLAS9 & IRAF splot & MOOG & $\log$ $\sigma$ (Fe) = 7.51, else \citet{Grevesse:1998p3102} & 91/22\\
\citet{Brugamyer:2011p3104} & 2.7m telescope at McDonald using the 2dcoude spectrograph and 9.2m Hobby-Eberly Telescope (HET) at McDonald using HIRES & 60\,000 (both) & 100-500 (both) & 3750-10200 /4090-7875 & ATLAS9 & IDL routine per \citet{Roederer10} & MOOG & their own & 65/22\\
\citet{Caffau:2011p3195} & 8.2m telescope at the Very Large Telescope (VLT) at ESO using the CRyogenic Infrared Echelle Spectrograph (CRIRES) & 60\,000 / 43000 & 670-800 & 5750-9310 & ATLAS12 & WIDTH9 per \citet{Kurucz1993} & WIDTH9 & their own & 46/17\\
\citet{Chen:2000p1857} & 2.16m telescope at Beijing Astronomical Observatory (BAO) with the Coude Echelle Spectrograph (CES) & 40\,000 & ~250 & 5600-8800 & MARCS & their own analysis & Uppsala EQWIDTH & differential analysis & 142/8\\
\citet{daSilva12} & Observatoire de Haute-Provence (OHP) using the ELODIE spectrograph & 42000 & $>$ 200 & 3895-6815 & ATLAS9 & ARES & MOOG & differential analysis & 72/12\\
\citet{DelgadoMena10} & ESO telescope using CORALIE & 115000 & 110 & 3800-6900 & ATLAS9 & ARES & MOOG & their own & 263/36\\
\citet{Ecuvillon:2004p2198,Ecuvillon:2006p8109} & 1.2m at the Euler Swiss Telescope (EST) using CORALIE, 2.2m telescope at ESO/MPI telescope using FEROS, the VLT Unit Telescope 2 (UT2) using the Ultraviolet and Visual Echelle Spectrograph (UVES), the 3.5m telescope at Telescopio Nazionale Galileo (TNG) using the SARG spectrograph, and the 4.2m telescope at William Herschel Telescope (WHT) using the Utrecht Echelle Spectrograph (UES) & R $>$ 50\,000 & 90-1350 & 3800-6800 & ATLAS9 & IRAF splot & MOOG & their own & 39/12\\
\citet{Edvardsson:1993p2124} & 1.4m telescope at ESO using CES and the 2.7m at McDonald using the 2dcoude spectrometer & 80\,000 / 50\,000 & 200-500 & varying ranges, not continuous & MARCS & Uppsala EQWIDTH & their own analysis & $\log$ $\sigma$(Fe) = 7.51 else \citet{Anders:1989p3165} & 32/2\\
\citet{Feltzing:1998p886} & 2.7m telescope at McDonald Observatory using the 2dcoude spectrometer & 10\,0000 & ~200 & 5200-8000 & MARCS & IRAF splot & per \citet{Edvardsson:1993p2124} & $\log$ $\sigma$(Fe) = 7.51 else \citet{Anders:1989p3165} & 43/4\\
\citet{Feltzing:2007p855} & 1.5m telescope at ESO using FEROS and 2.56m Nordic Optical Telescope (NOT) using the SOFIN spectrograph & 48000 / 80\,000 & ~200 & 5394, 5492, 6013, and 6016 & MARCS & their own analysis & their own analysis & their own & 176/36\\
\citet{Francois:1986p2312} & 1.4m telescope at ESO using the Coude Echelle Spectrometer and the 3.6m Canada France Hawaii (CFH) telescope using the coude focus & N/A & $>$ 150 & N/A & MARCS & N/A & ABON2 & \citet{Holweger:1979p3110} / Grevesse 1984 (S only) & 20 / 0\\
\citet{Fulbright:2000p2188} & 3m telescope at Lick Observatory using the Hamilton Echelle Spectrograph (HES), 2.1m telescope at McDonald using CASPEC, 10m telescope at Keck using HIRES, 3.6m telescope at ESO using Cassegrain Echelle Spectrograph (CASPEC) & 50\,000 & $>$ 100 & 3700-10075 & ATLAS9 & IRAF splot & MOOG & \citet{Anders:1989p3165} & 75/26\\
\citet{Gebran:2010p6243} & The OHP telescope using AURELIE and SOPHIE & 30\,000 or 60\,000 & ~200 & three ranges centered on 6160 , 5080 , and 5530 & ATLAS9 & spectral fitting & spectral fitting & \citet{Grevesse:1998p3102} & 20 / 0\\
\citet{Gilli:2006p2191} & 1.2m telescope at EST using CORALIE, 2.2m telescope at ESO/MPI using FEROS, the VLT/UT2 telescope using UVES, 3.5m telescope at TNG using SARG, and the 4.2m telescope at WHT using UES & 50\,000 / 48000 / 110\,000 / 57000 / 55000 & 150-350 & 3600-10100 & ATLAS9 & IRAF splot & MOOG & \citet{Anders:1989p3165} & 263/36\\
\citet{Gonzalez:2001p3080,Gonzalez:2007p1060} & 2.7m telescope at McDonald using the 2dcoude spectrometer and 4m Blanco Telescope at Cerro Tololo Inter-American Observatory (CTIO) & $>$ 59000 / 35000 & 195-620 & 3700-10\,000 / 5850-8950 & ATLAS9 & EQWIDTH & MOOG & their own & 64 / 11\\
\citet{GonzalezHernandez:2010p7714} & 3.6 m telescope at ESO using High Accuracy Radial Velocity Planet Searcher (HARPS) spectrograph, 8.2 m telescope at VLT/UT2 with UVES, and the 4.2m telescope at WHT using UES & 85000 & 800 & 3800-7950 & ATLAS9 & ARES & MOOG & differential analysis & 263 / 36\\
\citet{Gratton:2000p1209,Gratton:2003p1182} & VLT/UT2 telescope using UVES, 2.7m telescope at McDonald using the 2dcoude spectrometer, the NTT telescope using the EMMI spectrograph, 10m telescope at Keck using the HIRES spectrograph & 50\,000 / 80\,000 / 60\,000 / 50\,000 & 200 / 200 / 150 / $>$100 & 4700-7900 / 3700 - 9000 / N/A / 4400-9000 & ATLAS9 & their own analysis & EQWIDTH & their own & 46 / 17\\
\citet{Gustafsson:1999p8407} & 1.4m telescope at ESO using the CES & 65000 & 200-300 & Around 8727 & MARCS & spectral fitting & spectral fitting & differential analysis & 1 / 0 \\
\citet{Huang05} & 2.16m telescope at the National Astronomical Observatories with the CES & 40\,000 & 150-250 & 5600-9000 & ATLAS9 & ABONTEST & ABONTEST (via P. Magain) & differential analysis & 85/9\\
\citet{Jonsell:2005p1298} & 1.4m telescope at ESO using the CES & 60\,000 & 200 & 5670-5720 / 6120-6185 / 7750-7820 / 8710-8780 & MARCS & Uppsala EQWIDTH & EQWIDTH & \citet{Grevesse:1998p3102} / Asplund et al (2004) (O only) & 20 / 1\\
\citet{Kang11} & 1.8m telescope at the Bohyunsan Optical Astronomy Observatory (BOAO) using the Echelle Spectrograph & 30\,000-45000 & 150-250 & 3800-8800 & ATLAS9 & TAME via \citet{Kang12} & MOOG & their own & 82/7\\
\citet{Koch:2002p1283} & 1.4m telescope at ESO using the CES & 90\,000 & 50-340 & 4112-4148 & EAGLNT per \citet{Edvardsson:1993p2124} & spectral fitting & spectral fitting & their own & N/A\\
\citet{Korotin11} & 1.93m telescope at OHP using ELODIE & 42000 & 100-300 & 4400-6800 & STARSP code per \citet{Tsymbal96} & DECH20 per \citet{Galazutdinov92} & WIDTH9 & their own & 221/16\\
\citet{Laird:1985p1923} & 1m telescope at CTIO & N/A & N/A & 3300-5250 & \citet{Bell76} & their own analysis & MOOG & \citet{Holweger:1979p3110} & N/A\\
\citet{Luck:2005p1439} & 2.1m telescope at McDonald using the CASPEC & 60\,000 & $>$ 150 & 4840-7000 & MARCS75 & spectral fitting & spectral fitting & differential analysis & 450 / 25\\
\citet{Mashonkina11,Mashonkina:2007p1419} & 2.2m telescope at Calar Alto Observatory (CAO) using FOCES, 8m VLT/UT2 telescope using UVES & 60\,000 (both) & $>$ 200 & 4100-6800 & MAFAGS per \citet{Fuhrmann:1997p6740} & spectral fitting & spectral fitting & their own & N/A\\
\citet{Mishenina:2003p1368, Mishenina:2004p1360, Mishenina:2008p1380,Mishenina11} & 1.93m telescope at OHP using ELODIE & 42000 & 100-350 & 3850-6800 & ATLAS9 & WIDTH9 & N/A & differential analysis (2003, 2004, 2008), their own (2011) & N/A\\
\citet{Neves:2009p1804} & 3.6m telescope at ESO equipped with HARPS using CORALIE & 110\,000 & 70-2000 & 3800-6900 & ATLAS9 & ARES & MOOG & \citet{Anders:1989p3165} & 263/36\\
\citet{Nissen:2010p1903,Nissen11} & 8m VLT/UT2 telescope using UVES spectrograph and the NOT telescope using FIbre-fed Echelle Spectrograph (FIES) & 55000 / 40\,000 & 250-500 / 140-200 & 4800-6800 / 4000-7000 & MARCS & Uppsala EQWIDTH & EQWIDTH & differential analysis & 92/15 \\
\citet{Petigura:2011p3263} & 10m telescope at Keck using HIRES & 50\,000 & ~200 & 5000-6400 & ATLAS9 & spectral fitting & SME & \citet{Anders:1989p3165} & N/A\\
\citet{Ramirez:2007p1819,Ramirez:2009p1792} & 2.7m telescope at McDonald using the 2dcoude spectrograph, 1.52m telescope at ESO using FEROS, 9.2m telescope at HET at McDonald, and 8m VLT/UT2 telescope using UVES & 60\,000 / 45000 / 120\,000 / 80\,000 & 150-600 / 150-600 / ~300 / 300-50 & 4500-7800 (2007) and 3800-9125 (2009) & ATLAS9 & their own analysis & MOOG & differential analysis & 119/13\\
\citet{Reddy:2003p1354, Reddy:2006p1770} & 2.7 m telescope at McDonald using the 2dcoude spectrometer & 60\,000 & 100-200 & 3500-9000 & ATLAS9 & IRAF splot & MOOG & their own & 54/9\\
\citet{Shi:2004p2295} & 2.2m telescope at CAO using FOCES & 40\,000-60\,000 & 150-400 & 4000-9000 & MAFAGS & spectral fitting & spectral fitting & differential analysis & N/A\\
\citet{Takeda:2005p6195, Takeda:2007p1531} & 1.88m telescope at Okayama Astrophysical Observatory (OAO) using the High Dispersion Echelle Spectrograph (HIDES) & 70\,000 & 100-300 & 3900-8800 & ATLAS9 & WIDTH9 & SPSHOW (in the SPTOOL software developed by Y. Takeda, unpublished) & differential analysis (2005), \citet{Anders:1989p3165} (2007) & 160 / 20\\
\citet{Thevenin:1998p1499}) per \citet{Thevenin1999} & 9.2m telescope at HET at McDonald using HIRES & 60\,000 & 100-500 & 4090-7875 & \citet{Bell76} & Uppsala EQWIDTH & N/A & \citet{Holweger:1979p3110} & 2117 / 3445 (not all used) \\
\citet{Trevisan:2011p6253} & 1.52m telescope at ESO using FEROS & 48000 & 100 & 3560-9200 & MARCS & ARES & ABON2 via Spite 1967 (and improvements in the last 30yrs) & their own & 97/9\\
\citet{Valenti:2005p1491} & 10m telescope at Keck using HIRES, 4m telescope at AAT using UCLES, and 3m telescope at Lick using HES & 70\,000 & ~500 & 4830-6180 & ATLAS9 & spectral fitting & SME & \citet{Anders:1989p3165} & N/A\\
\citet{Wang11} & 1.88m telecope at OAO, 1.8m telescope at BOAO, and 2.16m telescope at Xinglong Station & 60\,000 & 150-230 & 4900-7600 & ATLAS9 & ABONTEST8 & ABONTEST8 & differential analysis & 81/8\\
\citet{Zhang:2006p6230} & 2.16m telescope at BAO with the CES & 37000 & 150 & 5600-8300 & ATLAS9 & ABONTEST8 & ABONTEST8 & \citet{Chen:2000p1857} & 64/11\\
\hline
\caption{Telescope/spectrograph information and the methods for determining abundances as given by the literature sources where $\ge$ 20 stars (see Table \ref{tab.cat}) were found in the Hypatia Catalog.}\label{tab.long}
\end{longtable}
\end{landscape}
\end{center}
\begin{center}\scriptsize
\begin{longtable}{|p{1.2cm}|p{0.7cm}|p{0.7cm}|p{0.7cm}|p{0.7cm}|p{0.7cm}|p{0.7cm}|p{0.7cm}|p{0.7cm}|p{0.7cm}|}
\hline
\multicolumn{1}{|l|}{\bf{Element}} &
\multicolumn{3}{l|}{\bf{ [Fe/H] = -0.4 dex}}&
\multicolumn{3}{l|}{\bf{ [Fe/H] = 0.0 dex}} &
\multicolumn{3}{l|}{\bf{ [Fe/H] = 0.4 dex}} \\
\bf{} & \bf{16\%} & \bf{Med} & \bf{84\%} & \bf{16\%} & \bf{Med} & \bf{84\%} & \bf{16\%} & \bf{Med} & \bf{84\%} \\
\hline
\hline
C & 0.16 & 0.25 & 0.34 & -0.18 & -0.07 & 0.09 & -0.24 & -0.15 & -0.07 \\
O & 0.12 & 0.28 & 0.40 & -0.34 & -0.09 & 0.06 & -0.48 & -0.36 & -0.25 \\
Na & -0.06 & 0.01 & 0.07 & -0.23 & -0.16 & 0.03 & -0.26 & -0.16 & -0.07 \\
Mg & 0.01 & 0.07 & 0.15 & -0.17 & -0.07 & 0.08 & -0.19 & -0.11 & -0.02 \\
Al & -0.15 & -0.09 & 0.03 & -0.17 & -0.08 & 0.02 & -0.17 & -0.12 & -0.07 \\
Si & 0.01 & 0.12 & 0.15 & -0.18 & -0.09 & 0.05 & -0.25 & -0.20 & -0.09 \\
K & N/A & 0.34 & N/A & N/A & -0.00 & N/A & N/A & N/A & N/A \\
Ca & 0.00 & 0.05 & 0.09 & -0.13 & -0.02 & 0.04 & -0.21 & -0.12 & -0.02 \\
Sc & -0.04 & 0.03 & 0.27 & -0.14 & -0.05 & 0.09 & -0.12 & -0.10 & 0.01 \\
ScII & 0.09 & 0.18 & 0.23 & -0.22 & -0.16 & 0.06 & -0.16 & -0.11 & -0.07 \\
Ti & -0.05 & 0.01 & 0.09 & -0.12 & -0.04 & 0.04 & -0.21 & -0.16 & -0.06 \\
TiII & N/A & 0.01 & N/A & N/A & -0.15 & N/A & N/A & -0.17 & N/A \\
V & -0.14 & -0.07 & 0.06 & -0.11 & -0.05 & 0.10 & -0.06 & 0.00 & 0.18 \\
Cr & -0.14 & -0.06 & 0.03 & -0.17 & -0.09 & -0.01 & -0.18 & -0.17 & -0.14 \\
CrII & -0.38 & -0.33 & -0.29 & -0.53 & -0.51 & -0.45 & -0.56 & -0.54 & -0.52 \\
Mn & -0.39 & -0.31 & -0.22 & -0.37 & -0.25 & -0.10 & -0.23 & -0.20 & -0.15 \\
Co & -0.06 & -0.02 & 0.06 & -0.18 & -0.08 & 0.01 & -0.06 & -0.02 & 0.11 \\
Ni & -0.06 & -0.04 & -0.00 & -0.22 & -0.15 & -0.01 & -0.22 & -0.16 & -0.09 \\
Cu & N/A & -0.21 & N/A & N/A & -0.12 & N/A & N/A & N/A & N/A \\
Zn & N/A & -0.15 & N/A & N/A & -0.13 & N/A & N/A & N/A & N/A \\
Sr & N/A & -0.27 & N/A & N/A & -0.18 & N/A & N/A & N/A & N/A \\
Y & N/A & -0.05 & N/A & N/A & 0.02 & N/A & N/A & N/A & N/A \\
YII & N/A & -0.12 & N/A & N/A & -0.03 & N/A & N/A & N/A & N/A \\
Zr & N/A & 0.08 & N/A & N/A & -0.06 & N/A & N/A & N/A & N/A \\
ZrII & N/A & -0.12 & N/A & N/A & 0.01 & N/A & N/A & N/A & N/A \\
BaII & -0.09 & -0.03 & 0.16 & -0.13 & -0.02 & 0.14 & N/A & N/A & N/A \\
Ce & N/A & -0.01 & N/A & N/A & 0.05 & N/A & N/A & N/A & N/A \\
Nd & N/A & 0.05 & N/A & N/A & 0.05 & N/A & N/A & N/A & N/A \\
Eu & N/A & 0.12 & N/A & N/A & -0.04 & N/A & N/A & N/A & N/A \\
\hline
\caption{Median and percentile 1$\sigma$ values for all elements with trend lines.}\label{tab.sigma}
\end{longtable}
\end{center}
\begin{table}[h]\scriptsize
\caption{
Number of Thin-Disk Stars in Hypatia with Spread Less than error bars
}
\begin{center}\tiny
\begin{tabular}{ c | c | c | c | c | c | c | c | c | c | c | c | c | c | c | c | c}
\hline
\hline
\bf{Li} & \bf{Be} & \bf{C} & \bf{N} & \bf{O} & \bf{Na} & \bf{Mg} & \bf{Al} & \bf{Si} & \bf{P} & \bf{S} & \bf{K} & \bf{Ca} & \bf{CaII} & \bf{Sc} & \bf{ScII} & \bf{Ti} \\
54 & 5 & 808 & 54 & 933 & 907 & 578 & 523 & 1098 & 1 & 162 & 139 & 657 & 8 & 381 & 386 & 1064 \\
\hline
\bf{TiII} & \bf{V} & \bf{VII} & \bf{Cr} & \bf{CrII} & \bf{Mn} & \bf{Fe} &\bf{Co} & \bf{Ni} & \bf{Cu} & \bf{Zn} & \bf{Sr} & \bf{SrII} & \bf{Y} & \bf{YII} & \bf{Zr} & \bf{ZrII} \\
227 & 564 & 13 & 483 & 329 & 512 & 1713 & 454 & 1035 & 242 & 203 & 119 & 25 & 191 & 210 & 96 & 188 \\
\hline
\bf{Mo} & \bf{Ru} & \bf{BaII} & \bf{La} & \bf{LaII} & \bf{Ce} & \bf{CeII} & \bf{Pr} & \bf{Nd} & \bf{Sm} & \bf{Eu} & \bf{EuII} & \bf{Gd} & \bf{Dy} & \bf{Hf} & \bf{Pb} & \\
48 & 19 & 301 & 73 & 39 & 237 & 5 & 75 & 203 & 29 & 119 & 34 & 62 & 17 & 73 & 17 &\\
\hline
\hline
\end{tabular}
\label{tab.skip}
\end{center}
\end{table}
\begin{table}[h]\scriptsize
\caption{
Scale Heights and 90\% Heights for Ensembles Above and Below [X/Fe] Gap
}
\label{scale.tab}
\begin{center}\tiny
\begin{tabular}{lcccc}
\hline
\bf{Element} & \bf{$h_a$ pc} & \bf{$h_b$ pc} & \bf{$h_{90a}$ pc} & \bf{$h_{90b}$ pc}\\
\hline
Mg & 35 & 22 & 76 & 39 \\
Si & 37 & 26 & 77 & 47 \\
S & 39 & 34 & 68 & 57 \\
Ca & 38 & 24 & 78 & 40 \\
Sc & 39 & 21 & 76 & 36 \\
Ti & 32 & 29 & 71 & 50 \\
Cr & 37 & 22 & 68 & 37 \\
Ni & 40 & 28 & 80 & 50 \\
\hline
\tablenotetext{a}{Standard composition values relative to \citet{Lodders:2009p3091} solar}
\end{tabular}
\end{center}
\end{table}
\begin{table}[h]\scriptsize
\caption{
Scale Heights and 90\% Heights for Ensembles Above and Below [X/Fe] Gap Without \citet{Neves:2009p1804}
}
\label{scale2.tab}
\begin{center}\tiny
\begin{tabular}{lcccc}
\hline
\bf{Element} & \bf{$h_a$ pc} & \bf{$h_b$ pc} & \bf{$h_{90a}$ pc} & \bf{$h_{90b}$ pc} \\
\hline
Mg & 35 & 24 & 76 & 42 \\
Si & 36 & 27 & 77 & 48 \\
S & 39 & 33 & 68 & 57 \\
Ca & 41 & 33 & 80 & 80 \\
Sc & 34 & 20 & 82 & 77 \\
Ti & 36 & 29 & 75 & 51 \\
Cr & 37 & 29 & 68 & 81 \\
Ni & 40 & 29 & 79 & 50 \\
\hline
\tablenotetext{a}{Standard composition values relative to \citet{Lodders:2009p3091} solar}
\end{tabular}
\end{center}
\end{table}
|
\section{Introduction}\label{introduction}
The study of statistical theories of fluctuating geometries
is important for a number of reasons. Regularized via appropriate lattices, so-called
Dynamical Triangulations (DT)\footnote{There are
other ways to provide lattice regularizations of
bosonic string theory, e.g. using hypercubic lattices \cite{dfj}.}),
they may serve as rigorous definitions of the path
integral of bosonic string theories \cite{DT},
or quantum gravity \cite{higherQG}. The DT formalism has
been immensely successful in the study of two-dimensional
quantum gravity coupled to conformal field theories with
central change $c \leq 1$, also known as non-critical string
theory or Liouville quantum gravity, but it has been less
successful serving as a regularization of a putative higher dimensional
quantum gravity theory \cite{firstorder}.
In an attempt to improve the situation
a modified lattice regularization, called the Causal Dynamical Triangulation (CDT),
was proposed in which a foliation structure is imposed on the
lattices representing space-time (which we here will assume
has Eucliean signature) \cite{al,ajl}). Such a foliation
structure is also imposed in the so-called Ho\v{r}ava-Lifshitz gravity
theory \cite{horava}. Some interesting results
related to higher dimensional quantum gravity have been obtained
using the CDT regularization (see \cite{physrep} for a review).
Here we will discuss the two-dimensional CDT theory, which
in principle should be simpler than the corresponding
two-dimensional DT theory. Indeed, the scaling limit of CDT not coupled to
matter is simple \cite{al,dgk}, and it can be shown to correspond
to two-dimensional Ho\v{r}ava-Lifshitz quantum gravity \cite{agsw}.
However, in
contrast to the case of two-dimensional DT,
it has been difficult to obtain solvable models of two-dimensional CDT coupled
to field theories. The only analytically solvable example of an explicit field theory
system coupled to gravity is provided by CDT coupled to gauge fields \cite{ai}, but
two-dimensional gauge field theories are mainly topological, so the
systems obtained are very simple from a matter perspective.
Computer simulations indicate that for unitary conformal field theories
with central charge $c \leq 1$ the coupling between matter and
geometry is weak \cite{aal,aalp} with
the critical exponents of both the matter
theories and the geometry apparently unchanged. This is in sharp contrast to
the DT situation, where both matter and geometric exponents are shifted
relative to the matter exponents in flat spacetime and the geometric
exponents in 2d Liouville gravity without matter. According
to the computer simulations the situation changes when the central
charge $c$ of the matter fields coupled to CDT is larger than one indicating
that the coupling to geometry then becomes strong \cite{aal1,krakow}.
Thus it was interesting and surprising when it was shown that
restricted dimer systems coupled to CDT could be solved
analytically and seemingly led to
a change of the critical exponents of the geometry \cite{atkin-zohren,aggs}.
It is well known that the hard dimer model on a regular two-dimensional
lattice exhibits critical behaviour for a certain negative value of
the fugacity, and that this critical system can be associated with
a (2,5) minimal non-unitary field theory having central
charge $c=-22/5$. In fact, it can be identified via the high temperature
expansion of the Ising model in an imaginary magnetic field with the Lee-Yang
edge singularity. In \cite{staudacher}
it was shown that a similar identification of a critical hard dimer model
with the Lee-Yang edge singularity can be made in the DT case
and the critical exponents can be calculated. One finds again a
non-trivial interaction between geometry and matter, but it is weaker
than the interaction between the unitary models and geometry. This is
in accordance with the expectation that when $c\to -\infty$ matter and gravity decouple.
Thus the change in critical
geometric properties found in \cite{atkin-zohren,aggs} when coupling the
CDT model to some classes of restricted dimers is puzzling: for unitary
models with $0<c\leq 1$ we have a weak coupling and no change
in critical geometric properties of the geometry, as mentioned above.
Naively we would expect the dimer systems at criticality to correspond
to non-unitary field theories with central charge $c <0$ and thus an
even weaker coupling, by analogy to the DT systems. This has motivated us
to take a closer look at the model proposed in \cite{atkin-zohren}
(hereafter called the AZ model). As we will report below the model is more subtle
than anticipated in \cite{atkin-zohren}.
The rest of this article is organized as follows. In Sec.\ 2 we define
a generalized AZ model and discuss its basic properties.
In Sec.\ 3 we consider the two-point function in the AZ model in a grand canonical
setting and we use it to calculate the global Hausdorff dimension
$d_H$. Sec.\ 4 addresses the calculation of the so-called
local Hausdorff dimension $d_h$ in a microcanonical setting. We find somewhat surprisingly
that $d_h=d_H=3$.
In Sec.\ 5 we show that this result is very special and probably not
representative for an unrestricted hard dimer model coupled to CDT.
We do this by analyzing in some detail the extended AZ model which allows more general dimer configurations while remaining solvable. The AZ model corresponds to one particular point in the phase boundary of this generalized model and we show that it is the only point at which the Hausdorff dimensions assume the value $3$, while the values at other points are either $d_H=3/2$ and $d_h=1$ or $d_H=d_h=2$. Sec.\ 6 contains a discussion of the results and arguments in favour of viewing $d_H=3/2$ as correct also
for the unrestricted dimer model coupled to CDT, i.e.\ for a $c=-22/5$
conformal field theory coupled to 2d Ho\v{r}ava-Lifshitz gravity. Finally, the Appendix discusses the general conditions responsible for the special features of the AZ model.
\section{A restricted dimer model, basic properties}\label{sec:2}
We consider an extension of the dimer system on random causal triangulations first introduced in \cite{atkin-zohren}.
A finite causal triangulation $T$ of the planar disc $\cal D$, is constructed as shown in Fig \ref{figCT}. $T$ is
the union of a central disc $\Sigma_0$ having \emph{central vertex} $v_0$ and
boundary circle $S_1$, and a sequence of annuli (or time slices) $\Sigma_k,\,k>1,$ such that $\Sigma_k$ is
bounded by circles $S_{k-1}$ and $S_k$. For $k\geq 1$, $\Sigma_k$ is triangulated by a circular array of triangles
each of which contains either one vertex in $S_{k-1}$ and two vertices in $S_k$, called a
\emph{backward directed triangle}, or two vertices in
$S_{k-1}$ and one vertex in $S_k$, called a \emph{forward directed triangle}. $\Sigma_0$
is triangulated by a sequence of triangles sharing the central vertex, which we define to be backward directed.
Edges contained in one $S_k$ are called \emph{horizontal edges}. By convention we adjoin a
forward directed triangle to each of the outermost horizontal edges (see Fig \ref{figCT}) so that all horizontal
edges are shared by a forward and a backward directed triangle.
We assume in the following that the edges/vertices in $S_k$ and triangles in $\Sigma_k$ are ordered clockwise, and it
is convenient to assume also that one of the edges emanating from the central vertex is marked.
\begin{figure}[t]
\centerline{\scalebox{0.4}{\includegraphics{cdt_disc_old.pdf}}}
\caption{A causal triangulation.}
\label{figCT}
\end{figure}
Given a vertex $v\neq v_0$ in $S_k$ we denote by $e(v)$ the horizontal edge in $T$
emanating in positive clockwise direction from $v$, by $\Delta(v)$ the forward directed triangle containing
$e(v)$ in its boundary, and by $f(v)$ the non-horizontal edge in $\Delta(v)$ emanating from $v$, see Fig \ref{bijection}.
Moreover, the \emph{forward degree} $\sigma_f(v)$ and the \emph{backward degree}
$\sigma_b(v)$ of $v$ are defined as the number of neighbours of $v$ in $S_{k+1}$ and $S_{k-1}$, respectively. Note
that $\sigma_f(v), \sigma_b(v) \geq 1$ and that $\sigma_f(v)=1$ if and only if $e(v)$ separates two forward directed
triangles.
Given a causal triangulation $T$, let $\tilde T$ denote its dual graph.
A \emph{restricted dimer configuration} $D$ on $\tilde T$ is a set of edges in $\tilde T$
fulfilling
\begin{enumerate}
\item[a)] no pair of edges in $D$ share a vertex in $\tilde T$;
\item[b)] edges in $\tilde T$ dual to edges in $T$ that separate two backward
directed triangles are not admissible in~$D$;
\item[c)] if $v\in S_k,\, k>0,$ is a vertex with $\sigma_f(v)=1$, then $f(v)$ is not dual to a dimer if either $\sigma_b(v)>1$ or if $\sigma_b(v)=1$ and the successor $w$ to $v$ in $S_k$ has $\sigma_b(w)>1$.
\end{enumerate}
Here, condition a) is the standard requirement specifying a dimer configuration, while b) and c) represent further technical
conditions allowing an exact solution by utilizing a mapping onto labelled trees as demonstrated below.
\begin{figure}
\centerline{\scalebox{0.6}{\includegraphics{bijection.pdf} } }
\caption{Labelling the edges of a forward directed triangle. The arrows indicate the clockwise direction on the
time slices.}
\label{bijection}
\end{figure}
The possible dimer types are illustrated in Fig \ref{figDimerTypes}. Dimers dual to horizontal edges we call type 1,
while those shared by a forward and a backward directed triangle in the same time slice we call types 2 and 2'
respectively depending on whether or not the backward triangle precedes the forward triangle w.r.t. clockwise ordering.
Type 3 dimers are those dual to edges shared by two forward directed triangles.
Finally we denote by $D_i$ the set of edges of type $i$ in $D$ so that
$D=D_1\cup D_2 \cup D_{2'}\cup D_3$.
The grand canonical ensemble we are interested in consists of elements $(T,D)$ specified by a causal triangulation
$T$ and an admissible dimer configuration $D$ on $\tilde T$. With the three types of dimers we associate
fugacities $\xi_1,\xi_2,\xi_{2'},\xi_3$ and define the partition function
\begin{equation}
Z(\xi_1,\xi_2,\xi_{2'},\xi_3;g) = \sum_{(T,D)} g^{|T|/2}\xi_1^{|D_1|}\xi_2^{|D_2|}\xi_{2'}^{|D_{2'}|}\xi_3^{|D_3|}\,,
\end{equation}
where $|A|$ denotes the number of elements in a set $A$. It is easy to show that $Z(\xi_1,\xi_2,\xi_{2'},\xi_3;g)$ is
well defined for any fixed values of $\xi_1,\xi_2,\xi_{2'},\xi_3$ provided that $|g|$ is sufficiently small.
It is straightforward to see \cite{atkin-zohren} that $Z$ is a function of $\xi_2+\xi_{2'}$ for fixed $g$ and $\xi_1,\xi_3$. Therefore, we shall
henceforth set $\xi_{2'}=0$, i.e. we further restrict dimer configurations such that $D_{2'}=\emptyset$, and drop $\xi_2'$
from the notation and write $\vec\xi = (\xi_1,\xi_2,\xi_3)$. Note that for $\xi_2=\xi_3=0$ we have
$Z(\xi_1,0,0;g)= Z(0,0,0;g(1+\xi_1))$, since the number of triangles in a causal triangulation equals twice the number
of horizontal edges and because horizontal dimers are mutually independent in the absence of non-horizontal dimers.
\begin{figure}[t]
\centerline{\scalebox{0.6}{\includegraphics{Dimer_types_v2.pdf}}}
\caption{Dimer Types. The arrow indicates the clockwise direction on a time slice.}
\label{figDimerTypes}
\end{figure}
In order to determine the analyticity properties of $Z$ we shall, as mentioned, exploit a correspondence between
admissible pairs $(T,D)$ and certain labelled trees which we now explain by slightly generalising a
construction given in \cite{atkin-zohren}.
From a causal triangulation $T$ one obtains a planar rooted
tree $\tau=\beta(T)$ in the following way:
\begin{enumerate}
\item[i)] delete all boundary
edges (those belonging to the outmost forward directed triangles);
\item[ii)] delete all horizontal edges $e(v)$ and all non-horizontal edges of the
form $f(v)$ for $v\in T$;
\item[iii)] attach a new root edge to $v_0$ such that the marked edge in $T$ is the rightmost
edge emanating from $v_0$ in $\beta(T)$.
\end{enumerate}
It was shown in \cite{DJW1} that $\beta$ yields a bijective correspondence between
causal triangulations with $2n$ triangles and rooted planar trees with n+2 vertices and root of order $1$. Note
that the vertices of $\beta(T)$ different from the root are also vertices of $T$ and that from now on when
referring to a tree we will denote the vertex next to the root by $v_0$ and call it the \emph{first vertex} of the tree. A dimer
configuration $D$ on $\tilde T$ induces the following labelling $\ell$ of the vertices of the tree $\tau=\beta(T)$:
\begin{enumerate}
\item[1)] if $e(v)$ is dual to a dimer in $D$, set $\ell(v)=1$;
\item[2)] if $f(v)$ is dual to a dimer in $D$ and $\sigma_f(v)>1$, set $\ell(v)=2$;
\item[3)] if $f(v)$ is dual to a dimer in $D$ and $\sigma_f(v)=1$, set $\ell(v)=3$;
\item[4)] the root is unlabelled;
\item[5)] otherwise, set $\ell(v)=0$.
\end{enumerate}
The restrictions imposed on $D$ are equivalent to the following constraints on $\ell$:
\begin{itemize}
\item[a)] if a leaf of $\beta(T)$ has label $3$ then its preceding neighbour at same height has label $0$ and
its successor at same height has label $0$ or $1$;
\item[b)] if a vertex in $\beta(T)$ that is not a leaf has label $2$, then its rightmost decendant does not have
label $1$;
\item[c)] a leaf of $\beta(T)$ that is the leftmost or the rightmost decendant of its predecessor does not have label $3$.
\end{itemize}
Noting that all edges in $T$ dual to dimers in $D$ are deleted when constructing $\beta(T)$, it is
straightforward to show that the correspondence between pairs $(T,D)$ and
pairs $(\tau,\ell)$ with $\ell(v_0(\tau))=0$ is bijective.
Note also that the number $\ell_i$ of vertices in $\beta(T)$ with label $i$ equals $|D_i|$ for $i=1,2,3$.
Consider now a labelled tree $(\tau,\ell)$ as above and assume the vertex $v_0$ next to the root has order $s+1$,
i.e. $\tau$ has $s$ branches $\tau_1,\ldots,\tau_s$ rooted at $v_0$. Since $\ell(v_0(\tau))=0$ the labellings of the $\tau_i$ induced by the labelling of $\tau$ are independent and the labelling of the first vertex is unrestricted.
However, the labellings of the different branches rooted at the vertex $v_0(\tau_i)$ may, depending on its label,
not be independent due to the constraints a) and b).
Now define the partition function for trees whose first vertex has label $i$ to be
\begin{equation}\label{W}
W_i(\vec\xi;g) = \sum_{(\tau,\ell):\ell(v_0)=i} g^{|\tau|}\xi_1^{\ell_1}\xi_2^{\ell_2}\xi_3^{\ell_3} ,\qquad i=0,1,2,
\end{equation}
where $|\tau|$ denotes the number of edges in $\tau$. Then decomposing trees into the root edge and their branches rooted at
$v_0$ shows that the $W_i$ satisfy the equations
\begin{equation} \label{j3} W_i=F_i(W_0,W_1,W_2; \vec\xi; g)\,,\end{equation}
where
\begin{eqnarray*}\label{Fdefinitions}
F_0&=&g\,(1-g\xi_3 W_0)H\\
F_1&=& g\,\xi_1(1-g\xi_3 W_0)H\\
F_2&=&g\,\xi_2\left(W_0+(1-g\xi_3 W_0)W_2\right)H
\end{eqnarray*}
with
\begin{equation}\label{H}
H = \left\{1-(1+g\xi_3) W_0 - W_1- (1-g\xi_3 W_0)W_2\right\}^{-1}\,.
\end{equation}
It follows from the discussion above that
\begin{equation}\label{ZW}
Z(\vec\xi;g) = g^{-1}W_0(\vec\xi;g) - 1\,,
\end{equation}
and that non-analytic behaviour of $Z(\vec\xi;g)$ as a function of $g$ at a critical point $g_c(\vec\xi)$, of the form
$$
Z_c - Z(\vec\xi;g) \sim (g_c-g)^\alpha\,,
$$
for some $0<\alpha<1$, occurs if and only if $W_0$ exhibits the same behaviour
\begin{equation}\label{gc}
W_{0c} - W_0(\vec\xi;g) \sim (g_c-g)^\alpha\,.
\end{equation}
Here and in the following the notation $a\sim b$ is used to indicate that there exist constants $c_1, c_2>0$ such
that $c_1\, a\leq b\leq c_2\, a$.
By eliminating $W_1, W_2$ from \eqref{j3} we obtain
\begin{equation}\label{W0}
(\xi_1+g\xi_3)\xi_2 W_0^3 -(1+\xi_1+\xi_2 +g\xi_3 + g^2\xi_2\xi_3) W_0^2 + (1+g\xi_2 + g^2\xi_3)W_0 -g =0
\end{equation}
which, for fixed $\vec\xi$, determines $W_0$ as the unique
root vanishing and analytic at $g=0$. For $\xi_1,\xi_2,\xi_3\geq 0$ the Taylor expansion of $W_0$ obtained from \eqref{W} has positive coefficients and hence its radius of convergence $g_c>0$ is a singularity of $W_0$. This is a property of $W_0$
that persists, as we shall see, in a larger range $S$ of couplings $\xi_1,\xi_2,\xi_3$ that are not necessarily positive. For generic $\vec\xi$ in this range, $W_{0c}$ is
a double root of \eqref{W0} at $g=g_c$
and $g_c$ is a square root branch point of $W_0$ as a function of $g$. If $\vec\xi\in S$ is such that $W_{0c}$ is a
triple root of \eqref{W0} at $g=g_c$ , i.e. $\alpha =1/3$ in \eqref{gc}, then
$(\vec\xi;g_c)$ is a multicritical point which we denote by $(\vec\xi_c;g_c)$.
The condition that $g_c$ be a double root is obtained by differentiating \eqref{W0} w.r.t $W_0$, so that
the critical coupling $g_c$ and the corresponding value $W_{0c}$ of $W_0$ satisfy
\begin{equation}\label{j9}
3(\xi_1+g_c\xi_3)\xi_2 (W_{0c})^2-2(1+\xi_1+\xi_2+g_c\xi_3+ g^2\xi_2\xi_3)W_{0c}+(1+g_c\xi_2 + g_c^2\xi_3)=0.
\end{equation}
Using this in \eqref{W0} yields
\begin{equation}\label{j10}
(1+\xi_1+\xi_2+g_c\xi_3+ g_c^2\xi_2\xi_3)(W_{0c})^2-2(1+g_c\xi_2 + g_c^2\xi_3)W_{0c} +3g_c=0.
\end{equation}
and $g_c$ and $W_{0c}$ are determined as functions of $\vec\xi$ by \rf{j9} and \rf{j10}.
Multicritical points additionally satisfy
\begin{equation}\label{j11}
3(\xi_{1c}+g_c\xi_{3c})\xi_{2c} W_{0c}-(1+\xi_{1c}+\xi_{2c}+g_c\xi_{3c}+ g_c^2\xi_{2c}\xi_{3c})=0\,.
\end{equation}
The existence of such multicritical points can be established by e.g.
setting $\xi_1=\xi$, $\xi_2=\kappa \xi$ and $\xi_3=0$, where $\kappa>0$. The value
$\kappa=2$ corresponds to the AZ model considered in \cite{atkin-zohren}. However
the results are universal for $\kappa>0$ and we will
denote all these models as AZ models, and explicitly
perform the calculations for $\kappa=1$ (except in footnote \ref{ftn1} where
we discuss the situation of an arbitrary real value of $\kappa$).
For $\kappa=1$ we use \eqref{j9},
\eqref{j10} and \eqref{j11} to obtain the equation
\begin{equation}\label{jjx}
\xi_c^3 + 24 \xi_c^2 + 3\xi_c -1 =0\,
\end{equation}
for the critical value of $\xi$.
This polynomial has one positive root and two negative roots and a
closer analysis shows that the largest negative
root $\xi_c \approx -0.278$ corresponds to a
multicritical point $(\xi_c,\xi_c,0;g_c)$,
while $g_c$ is a square root singularity
of $W_0$ for $\xi>\xi_c$ (for a discussion of the situation for a
general value of $\kappa$ see footnote \ref{ftn1}).
\section{The two-point function for $\xi_1=\xi_2=\xi, \xi_3=0$ }\label{sec:3}
We now consider the fractal behaviour of triangulations
and the corresponding labelled trees close to the critical point for the AZ model. The Hausdorff dimension in the grand canonical ensemble is determined
by the decay rate of the two point function \cite{book}
Concentrating first on trees define a marked
labelled tree to be a triple $(v,\tau,\ell)$ where $(\tau,\ell)$ is a labelled tree as above and $v$ is a vertex in
$\tau$ different from the root and the first vertex $v_0$. By $d(v)$ we denote the graph distance from the root to $v$.
The two-point function $\mathbb G(\xi,g;r)$ is defined by
$$
(\mathbb G(\xi, g;r))_{ij} = W_j^{-1}\sum_{(v,\tau,\ell):\ell(v_0)=i,\ell(v)=j,d(v)=r+1} g^{|\tau|}\xi^{\ell_1+\ell_2} \,,
$$
for $i,j\in\{0,1,2\}$ and $r\geq 1$.
A schematic illustration of the two-point function is shown in Fig.\ \ref{figj1}, from which it follows by standard arguments
and considerations similar to those leading to \eqref{j3} that
\begin{figure}[t]
\centerline{\scalebox{0.6}{\includegraphics{2_pt_function.pdf}}}
\caption{Graphical illustration of $\mathbb G(\xi,g;r)_{ij}$.}
\label{figj1}
\end{figure}
\begin{equation}\label{j4}
{\mathbb G}(\xi,g;r) = \mathbb T(\xi,g)^r\,,
\end{equation}
where
\begin{equation}
(\mathbb T(\xi,g))_{ij} =\frac{\partial F_i}{\partial W_j}
\end{equation}
which after some simplification gives
\begin{eqnarray}\label{j5}
\mathbb T (\xi,g)
&=& g^{-1}W_0^2
\begin{pmatrix}
1 & 1 & 1 \\
\xi &\xi & \xi \\
\xi(1-\xi W_0) & {\xi W_0}(1- \xi W_0)^{-1} & \xi (1 -\xi W_0) \end{pmatrix}\;.
\end{eqnarray}
Clearly, the matrix $\mathbb T$ has one eigenvalue $\lambda_3=0$. The two
other eigenvalues, $\lambda_1$ and $\lambda_2$, are solutions of the characteristic equation
\begin{equation}\label{j12}
\lambda^2 -\Big(-1+(1+\xi g)\frac{W_0}{g}\Big)\, \lambda +\frac{\xi^2W_0^3}{g}=0.
\end{equation}
which we rewrite as
\begin{equation}\label{j13}
\Big(\lambda-1\Big)\Big(\lambda -\frac{\xi^2W_0^3}{g}\Big) +
\lambda(1-\xi W_0) \,\frac{W_0}{g}\,\frac{\partial g}{\partial W_0}=0.
\end{equation}
On the critical line
$(W_{0c}(\xi),g_c(\xi)),\,\xi\geq \xi_c,$
the last term vanishes and
\begin{equation}\label{j14}
\lambda_1=1,~~~\lambda_2 = \frac{\xi^2W_{0c}(\xi)^3}{g_c(\xi)} ~~~~{\rm for}~~\xi \geq \xi_c.
\end{equation}
In particular, for $\xi=0$ we have $W_{0c}=1/2$, $g_c=1/4$ and $\lambda_2=0$. As $\xi$
decreases from zero to $\xi_c \approx -0.278$ we find that
$\lambda_2$ increases monotonically to
$\lambda_{2c} = 1$, where the value $1$ is a simple consequence of \eqref{W0} and \rf{j11}.
For fixed $\xi>\xi_c$ the two eigenvalues are real for $\Delta g= g-g_c$ small enough and the
dominant eigenvalue is $\lambda_1$ approaching $1$ as $g\to g_c$. Hence, we get in this case
$$
\mathbb G(\xi,g;r) = e^{-m(g)r + o(r)}
$$
as $r\to\infty$, where
$$
m(g) \sim \Delta \lambda_1 = 1-\lambda_1\,.
$$
At the critical line for $\xi > \xi_c$ we have $g'(W_{0})=0$ and
$g''(W_0)\neq 0$, i.e.\ $\alpha = 1/2$ in eq.\ \eqref{gc}. Thus
we obtain from \rf{j13}
\begin{equation}\label{j16}
\Delta\lambda_1 \sim \frac{\partial g}{\partial W_0}\sim |\Delta g|^\frac{1}{2}.
\end{equation}
Hence, the two-point functions decay exponentially with rate
$$
m(g) \sim |\Delta g|^{\frac 12}\,.
$$
This shows that for the labelled trees with $\xi>\xi_c$ the \emph{global Hausdorff dimension}, defined as the inverse of
the critical exponent of $m(g)$ (see e.g. \cite{book}), is $d_H=2$.
At the multicritical point $(\xi_c,g_c)$, on the other hand, we have $g'(W_0) = g''(W_0)=0$ and $\alpha = 1/3$ in \eqref{gc}.
Using this and \eqref{j11} gives
\begin{equation}\label{j18}
(1-\xi_c W_0) \frac{\partial g}{\partial W_0} = 3 \xi_c^2 (\Delta W_0)^2 + \xi_c\Delta g\,,
\end{equation}
where $\Delta W_0 = W_0-W_{0c}$. Inserting this expression into \eqref{j13} and setting $\Delta\lambda = 1-\lambda$ now gives\footnote{\label{ftn1} One can
repeat the calculations leading to \rf{j19} for the general assignment
$\xi_1=\xi$, $\xi_2= \kappa \xi$, $\xi_3=0$, $\kappa >0$, and
find that \rf{j19} is independent of $\kappa$. However
the values of $W_{0c}$, $g_c$ and $\xi_c$ depend on $\kappa$. What is
important is the existence of a multicritical point $\xi_c$.
For $\kappa=1$ this was ensured by eq.\ \rf{jjx}. The equation for a
general $\kappa$ is
$$
-b^3 \xi_c^3 +3(9-b^2)\xi_c^2 -3b\xi_c -1=0,~~~\kappa=b+2.
$$
For $\kappa >0$ the largest real negative root corresponds to
the multicritical point, precisely as in \rf{jjx}.
Interestingly, for the original AZ value
$b=0$ the equation simplifies to a trivial second order equation (it is the
point where the equation changes from having two negative and one
positive solution to one negative and two positive solutions, the third root
moving to $-\infty$ for $b\to 0^-$ and to $\infty$ for $b\to 0^+$). However,
this has no consequences for the discussion of multicriticality.
For $\kappa <0$ there is no negative real solution.}
\begin{equation}\label{j19}
(\Delta \lambda)^2 + \frac{3}{W_{0c}} \, \Delta \lambda \,\Delta W_0 +
\frac{3}{(W_{0c})^2}\, (\Delta W_0)^2 = O(\Delta W_0^3),
\end{equation}
and hence
\begin{equation}\label{j20}
\Delta \lambda = \frac{3}{2 W_{0c}} |\Delta W_0|(1\pm i /\sqrt{3}) + O(\Delta W_0^2).
\end{equation}
In particular, $\lambda_1 = \overline\lambda_2$ is complex for $g< g_c$ in this case and we conclude that
$\mathbb G(\xi,g;r)$ decays
exponentially (dressed with oscillating factors) with decay rate
$$
m(g) \sim \mbox{Re}\,\Delta\lambda \sim |\Delta W_0| \sim |\Delta g|^{\frac 13}\,.
$$
This yields the value $d_H=3$ for the global Hausdorff dimension at $\xi=\xi_c$.
Returning to the dimer model on causal dynamical triangulations the two-point function $G(\xi,g;r)$ is defined in
the same manner as for trees by marking a vertex $v$ at distance $d(r)$ from the central vertex of the triangulation $T$ and setting
$$
G(\xi,g;r) = \sum_{(v,T,D):d(v)=r} g^{|T|/2}\xi^{|D_1|+|D_2|}\,.
$$
Using the mapping $\beta$ between the dimer model and the labelled tree model we obtain
$$
G(\xi,g;r) = g^{-1}\sum_j \mathbb G_{0j}(\xi,g;r) W_j(g)\,.
$$
In particular, $G(\xi,g;r)$ has the same exponential decay rate as the two-point functions $\mathbb G(\xi,g;r)_{ij}$, and hence
the global Hausdorff dimension of the dimer model coincides with that of the labelled tree model, i.e. $d_H=2$ for $\xi>\xi_c$ and
$d_H=3$ for $\xi=\xi_c$.
\bigskip
\section{The infinite size limit}\label{sec:4}
In this section we consider an alternative notion of Hausdorff dimension for the labelled tree model by considering only critical
trees, i.e. we shall evaluate the infinite size limit first and express the Hausdorff dimension in terms of volume growth on
infinite trees. In order to make this precise, let us introduce the finite size partition functions $W_{iN}(\vec\xi)$ by restricting the
sum in \eqref{W} to trees $\tau$ of fixed size $N$, so that
$$
W_i(\vec\xi;g) = \sum_N g^N W_{iN}(\vec\xi)\,.
$$
The distributions $\mu_{iN}$ of labelled trees of fixed size $N$ are obtained by normalizing the weights defining $W_{iN}$, i.e.
$$
\mu_{iN}(\tau,\ell) = \frac{1}{W_{iN}(\vec\xi)} \xi_1^{\ell_1}\xi_2^{\ell_2}\xi_3^{\ell_3}\,.
$$
Obviously, $\mu_{0N}$ is non-negative whenever $\xi_1,\xi_2,\xi_3\geq 0$ and hence defines a probability distribution. For $\xi_1=\xi_2=\xi$ and $\xi_3=0$ it is not difficult to see that this even holds true for $\xi\geq -\frac 14$, but not for $\xi_c\leq \xi< -\frac 14$. Similar remarks apply to $\mu_{iN}$ up
to a sign factor. Our aim is to consider limits of the expectations $\langle\cdot\rangle_{iN}$ with respect to the (signed) distributions $\mu_{iN}$ as $N\to\infty$, for arbitrary values of $\vec\xi\in S$.
As a consequence of \eqref{gc} and standard transfer theorems (see e.g. \cite{FS:2009}) the following asymptotic behaviour of $W_{iN}(\vec\xi)$ for large $N$ holds:
\begin{eqnarray}
W_{iN}(\vec\xi) &=& \Omega_i N^{-3/2} g_c(\vec\xi)^{-N}(1+\mbox{O}(\tfrac 1N))\quad\mbox{if $\alpha=1/2$}\,,\label{asymp1}\\
W_{iN}(\vec\xi) &=& \Omega_i N^{-4/3} g_c(\vec\xi)^{-N}(1+\mbox{O}(\tfrac 1N))\quad\mbox{if $\alpha=1/3$}\,,\label{asymp2}
\end{eqnarray}
where the constants $\Omega_i$ depend on $\vec\xi$. We note that the relations \eqref{j3} imply
\begin{eqnarray}
\Omega_1 &=& \xi_1\Omega_0 \label{Omeg1}\\
\Omega_2 &=& \xi_2 W_{0c}\frac{2-\xi_2 W_{0c}-g_c\xi_3W_{0c}}{(1-\xi_2 W_{0c})^2(1-g_c\xi_3W_c)^2}\,\Omega_0\,. \label{Omeg2}
\end{eqnarray}
Using \eqref{asymp1} and \eqref{asymp2} it follows by a straight-forward generalization of arguments given in \cite{BD,DJW1} that
for any local quantity $A(\tau,\ell)$ depending only on the structure of $(\tau,\ell)$ within a finite distance $R$ from the root of $\tau$, such
as the volume of the ball $B_R(\tau)$ of radius $R$ centered at the root, the limiting expectation values
$$\langle A\rangle_i = \lim_{N\to\infty}\langle A\rangle_{iN}$$
exist.
\begin{figure}
\centerline{{\includegraphics[scale=0.6,angle=-90]{grt.pdf}}}
\caption{Finite trees attached to the first few vertices on the spine of an infinite tree.}
\label{SpineTree}
\end{figure}
We next briefly describe how to calculate the limiting expectation values $\langle A\rangle_i$ in terms of infinite labelled trees, further details can be found in \cite{BD,DJW1}. Here an infinite labelled tree means an infinite rooted planar tree with root of order $1$ with a
labelling respecting the same conditions a)- c) as previously. Moreover, only trees with a single spine, that is an infinite
self-avoiding path starting at the root, contribute to $\langle A\rangle_i$, see Fig. \ref{SpineTree}.
The spine vertices of an infinite labelled
tree $L$ will be denoted by $r, u_1, u_2, u_3,\dots$, ordered by increasing distance from the root $r$. Thus $L$ is obtained
by grafting finite labelled trees with root of order $1$, called branches, at the spine
vertices $u_i$ on both sides of the spine. Considering only the finite part $r,u_1, u_2,\dots, u_N$ of the spine one of the branches
rooted at $u_N$ is infinite and all other branches are finite. We denote by $L(N)$ the finite labelled tree obtained by removing
the infinite branch at $u_N$ except the edge $u_Nu_{N+1}$ and consider $L(N)$ as a finite labelled tree with a finite
spine $r, u_1, u_2,\dots, u_{N+1}$ both of whose end vertices have order $1$. By ${\cal A}(L_0)$ we denote the set of all
infinite labelled trees such that $L(N)$ equals a fixed finite labelled tree $L_0$ with a distinguished spine $r, u_1, u_2,\dots, u_{N+1}$.
With this notation the limiting weight of the set ${\cal A}(L_0)$ equals
\begin{equation}\label{mui}
\mu_i({\cal A}(L_0)) = \Omega_i^{-1}\Omega_{\ell_{N+1}} \rho_i(L_0)\,,
\end{equation}
where $\rho_i(L_0)$ is the grand canonical weight of $L_0=(\tau_0,\ell_0)$ at the critical point $(\vec\xi;g_c)$ given by
\begin{equation}\label{rhoi}
\rho_i(\tau_0,\ell_0) = \delta_{\ell_0(u_1),i} g_c^{|\tau_0|-1} \xi_1^{\ell_{01}-\delta_{\ell_0(u_{N+1}),1}}\xi_2^{\ell_{02}-\delta_{\ell_0(u_{N+1}),2}}\xi_3^{\ell_{03}} \,.
\end{equation}
The information contained in \eqref{mui} and \eqref{rhoi} suffices to calculate $\langle A\rangle_i$ for any local quantity $A$.
We now proceed to calculate $\langle |B_R|\rangle_i$, where $|B_R(L)|$ denotes the size of $B_R(L)$, i.e. the number of edges
in $\tau$ whose vertices are at graph distance at most $R$ from the root. The \emph{local Hausdorff dimension} $d_h$ of the
random tree defined by $\mu_i$ is defined by
\begin{equation}\label{haus2}
\langle\; |B_R|\;\rangle_i \; \sim \; R^{d_h}
\end{equation}
as $R\to\infty$. The purpose of the next two subsections is to evaluate $d_h$ in the case $\xi_3=0$ and demonstrate that its value
coincides with $d_H$ as found in Section 3
\subsection{Volume of a finite tree for $\xi_1=\xi_2=\xi$, $\xi_3=0$}
We let $H_i^R$ denote the (unnormalized) expectation value of the number
of vertices $h_R(\tau)$ at distance $R$ from the root
of a finite tree $\tau$ with label $i$ on its vertex $v_0$ at the critical value $g_c(\xi)$ of the coupling $g$, that is
\begin{eqnarray}
H_i^R = \sum_{(\tau;\ell):\ell(v_0)=i} h_R(\tau) g_c^{|\tau|} \xi^{\ell_1+\ell_2}\,.
\end{eqnarray}
Applying arguments similar to those leading to \eqref{j3} one obtains, for $i=0,1,2$,
\begin{eqnarray}
H_i^1 &=& W_{ic}\,,\nonumber\\
H_i^R&=&\sum_{j=0}^2\mathbb T(\xi,g_c(\xi))_{ij} H_j^{R-1}\,,\quad R\geq 2,\label{H3}
\end{eqnarray}
where $\mathbb T$ is given by \eqref{j5}. $\mathbb T_c = {\mathbb T}(\xi,g_c(\xi))$ has eigenvalues
\begin{equation} \lambda_0=0, \quad
\lambda_1=1,\quad \lambda_2 = \frac{\xi^2W_{0c}(\xi)^3}{g_c(\xi)}\,,\label{Tevals}
\end{equation}
and right eigenvectors corresponding to the non-zero eigenvalues
\begin{equation}
\mathbf e^{(1)}=\mathbf M= \frac{W_{0c}^2}{g_c}\begin{pmatrix} 1\\ \xi \\ {g_c}W_{0c}^{-2}-1-\xi \end{pmatrix},\; \mathbf e^{(2)}= \frac{W_{0c}^2}{g_c}\begin{pmatrix} 1\\ \xi \\ \lambda_2{g_c}W_{0c}^{-2}-1-\xi \end{pmatrix}.\label{Tevecs}
\end{equation}
Here
$$
\mathbf M = \Omega^{-1}(\Omega_0,\Omega_1,\Omega_2)\,,
$$
where
$$
\Omega = \sum_{i=0}^2 \Omega_i \,,
$$
such that $\sum_i M_i=1$.
There are now two cases to consider:
\begin{itemize}
\item[\underline{$\xi >\xi_c$}]
The eigenvalue $\lambda_2<1$, $\mathbb T_c$ is diagonalizable and it is straightforward to show that
\begin{eqnarray}\label{Hanswer} {\mathbf H}^R=(1-\lambda_2)^{-1}
\left[
(1-(\lambda_2+1)g_cW_{0c}^{-1})\,{\mathbf M}
+ \lambda_2^{R-1} (2g_cW_{0c}^{-1}-1)\,{\mathbf e}^{(2)}
\right]. \end{eqnarray}
\item[\underline{$\xi =\xi_c$}]
The eigenvalue $\lambda_2=1$ and we see from \eqref{Tevecs} that its eigenvector coincides with $\mathbf M$ so $\mathbb T_c$ has
non-trivial Jordan normal form and a new vector
\begin{equation} \boldsymbol\varepsilon=\begin{pmatrix} 0\\0\\1\end{pmatrix}\end{equation}
emerges satisfying
\begin{equation}\label{Jshift} \mathbb T_c\, \boldsymbol\varepsilon=\mathbf M+\boldsymbol\varepsilon.\end{equation}
Setting
$$
{\mathbf W} = (W_0,W_1,W_2)
$$
and noting that
\begin{equation}\label{Wcreln}
{\mathbf W}_c=g_cW_{0c}^{-1}{\mathbf M}+(1-2g_cW_{0c}^{-1}){ \boldsymbol \varepsilon}
\end{equation}
then gives
\begin{equation}\label{HRc}
\mathbf H^R = \mathbf W_{c} + (R-1)(1-2g_cW_{0c}^{-1})\mathbf M\,,\quad R\ge 1\,.
\end{equation}
\end{itemize}
\subsection{Volume of an infinite tree for $\xi_1=\xi_2=\xi$, $\xi_3=0$}
We let $K_i^R$ denote the (unnormalized) expectation value with respect to the measure $\mu_i$ of the number $k_R(\tau)$ of vertices at distance
$R$ from the root of an infinite tree $\tau$ up to a normalization factor. Specifically,
$$
K_i^R = \Omega^{-1}\,\Omega_i\, \langle k_R \rangle_i\,.
$$
By decomposing the tree into its branches at the vertex $u_1$ next to the root one finds that $ K_i^R$ satisfies
\begin{equation} K_i^{R} =\frac{\partial F_i}{\partial W_j} K_j^{R-1} + M_j\frac{\partial^2 F_i}{\partial W_j\partial W_k} H_k^{R-1}\end{equation}
or
\begin{equation} {\mathbf K}^{R} = \mathbb T_c {\mathbf K}^{R-1}+\Gamma {\mathbf H}^{R-1}\label{Kprop}\end{equation}
where
\begin{equation} \Gamma_{ik} = M_j\frac{\partial^2 F_i}{\partial W_j\partial W_k} = M_j\Lambda_{i,jk}. \end{equation}
The first term in \eqref{Kprop} is the contribution of the infinite branch and the second term that of the finite branches. As each tree has only a single vertex at height 1,
\begin{equation} {\mathbf K}^{1}={\mathbf M}.\end{equation}
The equation \eqref{Kprop} is easily iterated to get
\begin{eqnarray} {\mathbf K}^{R} &=& {\mathbf M} + \sum_{\ell=1}^{R-1} {\mathbb T}_c^{\ell-1}\,\Gamma \,
{\mathbf H}^{R-\ell}.
\label{Kformula}\end{eqnarray}
There are again two cases to consider:
\begin{itemize}
\item[\underline{$\xi >\xi_c$}] Combining \eqref{Hanswer} and \eqref{Kformula} and noting that
\begin{equation}\label{GM}
\Gamma\, {\mathbf M}= 2g_c^{-1}W_{0c} \,{\mathbf M}\end{equation}
gives
\begin{eqnarray} {\mathbf K}^{R}
&=& 2\,\frac{(W_{0c}g_c^{-1}-\lambda_2-1)}{1-\lambda_2} \,R\,\mathbf M + O(1)
\end{eqnarray}
from which we get
\begin{equation}
\langle |B_R|\rangle_i = \Omega_i^{-1}\, \Omega\,\sum_{n=1}^R K_i^n = \Omega\,
\frac{(W_{0c}g_c^{-1}-\lambda_2-1)}{1-\lambda_2}\,R^2 +O(R).\label{Bresult}\end{equation}
It is straightforward to check that the coefficient of the $R^2$ term is positive for all $\xi>\xi_c$ so we have shown
that $d_h = 2$ in this regime. Note also that the coefficient diverges at $\xi=\xi_c$, where $\lambda_2\to 1$, indicating
that $d_h$ changes there.
\item[\underline{$\xi =\xi_c$}] Combining \eqref{HRc} and \eqref{Kformula} we have
\begin{eqnarray}\label{basic} {\mathbf K}^{R} = {\mathbf M} &+& \sum_{\ell=1}^{R-1} {\mathbb T}_c^{\ell-1}\,\Gamma \left(
(1-2g_cW_{0c}^{-1}) ({ \boldsymbol \varepsilon}-{\mathbf M}) + g_cW_{0c}^{-1}{\mathbf M}\right)\nonumber\\
&+& \sum_{\ell=1}^{R-1} {\mathbb T}_c^{\ell-1}\,\Gamma \left( (R-\ell)(1-2g_cW_{0c}^{-1}){\mathbf M} \right)
\end{eqnarray}
It is straightforward to check that at the tricritical point
\begin{equation}
\Gamma\, { \boldsymbol\varepsilon} = 2g_c^{-1}W_{0c}\left({\mathbf M} +\frac{1}{2} { \boldsymbol\varepsilon}\right).
\end{equation}
Using this identity and \eqref{Jshift} then gives
\begin{equation} {\mathbf K}^{R} =3 \left(\frac{W_{0c}}{2g_c}-1\right) R^2\,
{\mathbf M} +O(R).\end{equation}
It is worth noting that one might have supposed from \eqref{Jshift}, \eqref{HRc} and \eqref {Kformula}
that ${\mathbf K}^{R}$ would be $O(R^3)$;
however the coefficient of this leading term vanishes as a
consequence of the tri-criticality condition. It follows now that
\begin{equation}
\langle |B_R|\rangle_i = \Omega_i^{-1}\, \Omega\, \sum_{n=1}^R K_i^n = \Omega \left(\frac{W_{0c}}{2g_c}-1\right) R^3\, +O(R^2)
\end{equation}
The coefficient of the $R^3$ term evaluates to a
positive number so we have shown that $d_h = 3 $ at $\xi=\xi_c$ in the AZ model.
\end{itemize}
\section{The extended model $\xi_3\ne0$}\label{sec:5}
In \cite{atkin-zohren}
it was argued that the AZ model with the CDT coupled to a reduced set of dimers is not likely to differ significantly from the full CDT-with-dimers system (hereafter called CDT-D). We claim here that this is probably not correct by considering what happens when $\xi_3\ne0$; this perturbation is arguably closer to the CDT-D model as it incorporates more of the possible dimer types than the AZ model.
In the most general case $\mathbb T$ is given by
\begin{equation}\label{Tgen}
\mathbb T (\xi;g)= H
\begin{pmatrix}
g(1+g\xi_3W_1)H & W_0 & W_0(1-g\xi_3W_0) \\
g\xi_1(1+g\xi_3W_1)H & W_1& W_1(1-g\xi_3W_0) \\
g\xi_2(1-W_1( 1-g\xi_3W_2 ))H & W_2 & \xi_2 W_0(1-W_1 -g\xi_3 W_0)\end{pmatrix}\;.
\end{equation}
It is straightforward although tedious to show that on the critical surface $g_c(\xi_1,\xi_2,\xi_3)$
\begin{equation} \mathbf M = \frac{W_{0c}^2}{g_c}\left(1+\xi_3(1+\xi_1)W_0^3(2-g\xi_3W_0)\right)^{-1}\begin{pmatrix} 1\\ \xi_1 \\ {g_c}W_{0c}^{-2}-(1+\xi_1) (1-g\xi_3W_0)^2
\end{pmatrix}\end{equation}
is always a right eigenvector with eigenvalue 1 and that the other eigenvalues are 0 (corresponding to the fact that $F_1=\xi_1 F_0$) and
\begin{equation}\label{lambda2gen} \lambda_2= \frac{\xi_1\xi_2 W_0^3}{g(1-g\xi_3W_0)^2} = g\xi_1\xi_2 W_0 H^2\,, \end{equation}
where we have used
$$
H = \frac{W_0}{g(1-g\xi_3 W_0)}\,.
$$
For definiteness we will first discuss the model with $\xi_1=\xi_2=\xi_3=\xi$ which, at least naively, is the closest we can get to CDT-D.
For $\xi >\xi_{c} \approx -0.228$ we find that $g_c$ is a square root singularity of $W_0$, so $\alpha=\frac{1}{2}$, and
$\lambda_2<1$ at $g=g_c$.
Consequently $m(g)\sim\vert\Delta g\vert^\frac{1}{2}$ and $d_H=2$ following the discussion of Section \ref{sec:3}. The infinite graph calculation of the local Hausdorff dimension follows the same lines as the $\xi>\xi_{c}$ case in Section \ref{sec:4} leading to $d_h=2$. In this region of parameter space, where the dimer system is not critical, the model has exactly the same properties as the AZ model.
However at $\xi =\xi_{c}$ there is a tricritical point at which $\mathbb T_c$
is diagonalisable, $\lambda_2\approx 0.445 <1$ and $m(g)\sim \vert\Delta g\vert^\frac{2}{3}$ (the absence of a $\vert\Delta g\vert^\frac{1}{3}$ term is a consequence of the tricriticality condition).
Thus $d_H=\frac{3}{2}$ but, as shown in the Appendix, $d_h=1$.
It is interesting to compare this result with the simpler
multicritical tree model of rooted binary trees with dimers
placed on the edges, including the root edge, in such a way
that no more than one dimer can end at any vertex
\cite{aggs}. Letting $W_1$ and $W_0$ be the partition functions for trees with and without a dimer on the root edge respectively we see that they satisfy equations of the same form as \eqref{j3} but with
\begin{eqnarray} F_0&=&g(W_0^2+2W_0W_1+1)\nonumber\\
F_1&=&g\xi(W_0^2+1).
\end{eqnarray}
This model also has a tricritical point with $\xi_c=-\frac{4}{27}$, exponent $\alpha=\frac{1}{3}$ and $\lambda_{2c}<1$. One finds that
\begin{equation}\label{j17}
\Delta\lambda_1 \sim \frac{\partial g}{\partial W_0}\sim (\Delta W_0)^2 \sim |\Delta g|^{2/3}.
\end{equation}
which implies that $d_H=3/2$ for $\xi=\xi_c$. On the other hand, using the results of the Appendix, $d_h=1$ as there is once
again only one unit eigenvalue of $\mathbb T_c$. The $\xi_1=\xi_2=\xi_3$ line of our model thus exhibits exactly the same behaviour
as a standard multi-critical tree model.
\begin{figure}[t]
\centerline{\scalebox{0.6}{\includegraphics{PhaseDiagram.pdf}}}
\caption{The phase diagram in the $(\xi_1=\xi_2=\xi,\, \xi_3)$ plane.
The solid line is the line of cubic degeneracies in $W_0$.
For $\xi_3 >0$ the physical region $S$ defined in Sec.\ \ref{sec:2}
lies to the right of the long dashed line, which is the line
where $\lambda_2 =1$. For $\xi_3 \leq 0$ the first part of the
boundary of $S$ is the lower part of the solid line of
cubic degeneracies in $W_0$. The dotted line makes up the rest of the lower
part of the boundary of $S$. Along this part $W_0$ is only quadratic
degenerate.}
\label{figPhaseDiagram}
\end{figure}
These calculations appear to show that the degenerate tri-critical point with $d_h=3$ found in the AZ model is very special and
not at all characteristic of CDT dimer models in general. It is instructive to examine the phase diagram in the
$(\xi_1=\xi_2=\xi, \,\xi_3)$ plane, see Fig \ref{figPhaseDiagram}. There is a line of cubic degeneracies in $W_0$ that takes in
the point $(-0.278\ldots, 0)$ and extends both above and below the $\xi$ axis. Using \eqref{lambda2gen} and the identity
$$
g^2\xi^2H^3 - g^2(1+\xi)\xi_3 H^3 -1=0
$$
which holds at tricritical points as a consequence of \eqref{j9},\eqref{j10} and \eqref{j11}
it follows that
$$
\lambda_2 = 1 + g_c \xi_3 (1+\xi - \xi^2 W_0) H^3\,.
$$
Since the expression in parenthesis, $H$ and $g_c$ are all positive it follows that on the tricritical line $\lambda_2 <1$
for $\xi_3<0$ but that $\lambda_2 >1$ for $\xi_3>0$. The latter behaviour is a little strange; it would in fact be a contradiction
for a purely real eigenvalue of $\mathbb T$ to go through 1 \emph{before} criticality is reached
(see the Appendix for example). Closer inspection shows that at small $g\ll g_c$ the eigenvalues are complex and as $g$
increases they flow as shown in Fig \ref{Tevalflow}. However exponential growth of the two-point function is a symptom that the
series in $\xi_i$ and $g<g_c$ for $Z$ is not absolutely convergent and the effect of the negative weights is
sufficiently strong that the conventional statistical
mechanical interpretation of the model fails. We conclude that the physical region for $\xi_3>0$ extends only as far as the
line where $\lambda_2 =1$ at $g_c$. It
can be checked that inside the region and along this line there are only quadratic degeneracies in $W_0$ and that $\mathbb T_c$
is diagonalisable so $d_h=d_H=2$. On the other hand for $\xi_3<0$ there is a genuine line of tricriticality which includes
the $\xi=\xi_3$ point analysed above and ends at $(-0.162\ldots,-0.582\ldots)$. Beyond this point the tricriticality
disappears and even on the boundary of the physical region $\alpha=\frac{1}{2}$ and $d_h=d_H=2$.
\begin{figure}
\centerline{\scalebox{0.6}{\includegraphics{T_eval_flow.pdf}}}
\caption{The flow in the complex plane of the non-zero eigenvalues of $\mathbb T$; the arrow heads show how they move as $g$ increases; $\lambda_1$ ends at 1, $\lambda_2$ at a value $>1$. }
\label{Tevalflow}
\end{figure}
\section{Concluding remarks}
At the critical value of the dimer fugacity we expect that CDT-D, the full dimer model on CDT, represents a lattice regularization of projectable Ho\v{r}ava-Lifshitz quantum gravity
coupled to a (2,5) minimal conformal field theory.
In the absence of a solution to CDT-D we have obtained
the solution of a restricted dimer model and mapped out its phase diagram. In particular,
we have seen that the geometric features of the AZ model \cite{atkin-zohren} are very
special and not robust under perturbations. At generic points of the phase
boundary we have found the values of the Hausdorff dimensions are either $d_H=3/2$ and
$d_h=1$, coinciding with the values for the simplest multicritical tree
\cite{adj,book}, or $d_H=d_h=2$.
One may speculate on the implications of our results for the CDT-D model.
While we do not have any rigorous results in this direction it is
worth noting that
the full dimer model on a generalized CDT \cite{GCDT} has been solved
exactly in \cite{aggs} using a matrix model representation yielding
the value $d_H=3/2$.
The generalized causal triangulations of this model
can be defined combinatorially
\cite{ab} or by using a special scaling limit of matrix models \cite{GCDT}.
This slightly more general set of triangulations has many of the
characteristics of CDT, e.g.\ $d_H=d_h=2$. Hence it is tempting on
the basis of this result
to conjecture that $d_H=3/2$ and $d_h=1$ are indeed the correct
values for the full dimer model on a CDT.
It is natural to extend the above considerations further. One can
define multicritical generalized CDT models \cite{az-multi}, which
most likely correspond to specific fine-tuned scaling limits of
matrix models, generalizing the considerations in \cite{aggs}.
Recall that the standard multicritical matrix models from DT provide
representations of 2d Euclidean quantum gravity coupled to
certain conformal field theories. They also have the interpretation
of (increasingly complicated) fine-tuned multi-dimers systems on the
DT-set of random graphs. Thus it is possible that the multicritical
behavior found in \cite{az-multi} represents the effect of
fine-tuned multi-dimer models on the generalized CDT-set of random
graphs, and has the continuum interpretation of certain conformal
field theories coupled to 2d Ho\v{r}ava-Lifshitz gravity.
This leave us with the interpretation of the AZ model. Although this
model is special, it is not {\it that} special. As we have seen there
is a least a one-parameter set of coupling constants leading to the same
scaling. Thus we believe there should also be a
continuum interpretation of this class of models.
\subsection* {\ Acknowledgements.}
J.A. acknowledges support from the ERC Advanced Grant 291092
``Exploring the Quantum Universe'' (EQU) as well as from
the Free Danish Research Council grant ``Quantum gravity and the
role of black holes.'' In addition JA was supported
in part by the Perimeter Institute of Theoretical Physics.
Research at Perimeter Institute is supported by the Government of Canada
through Industry Canada and by the Province of Ontario through the
Ministry of Economic Development \& Innovation.
J.A. and B.D. acknowledge support from NordForsk researcher
network Random Geometry (grant no. 33000)".
\section*{Appendix: \boldmath{$d_h$} and the multi-critical condition }
We discuss here how the condition for multi-criticality affects $d_h$ in general. We consider a set of generalised trees with first vertex label $\ell(v_0)=i$ whose partition functions
$W_i$ satisfy
\begin{equation} W_i = F_i(\mathbf W;\vec\xi;g),\qquad i=1\ldots N.\end{equation}
The approach to criticality in the grand canonical ensemble (GCE) is governed by (repeated indices are summed over)
\begin{equation} \delta W_i = \delta g \frac{\partial F_i}{\partial g}+ \mathbb T_{ij}\delta W_j + \frac{1}{2} \Lambda_{i,jk}\,\delta W_j \delta W_k+{\mathrm {h.o.t.}}\end{equation}
where
\begin{eqnarray} \mathbb T_{ij}= \frac{\partial F_i}{\partial W_j} ,\quad
\Lambda_{i,jk}=\frac{\partial^2F_i}{\partial W_j\partial W_k},\quad
(\mathbf\Lambda_{jk})_i=\Lambda_{i,jk}.
\end{eqnarray}
Re-arranging
\begin{equation} ((1-\mathbb T)\delta W)_i = W_i\delta g + \frac{1}{2} \Lambda_{i,jk}\,\delta W_j \delta W_k+{\mathrm {h.o.t.}}\label{GCE}\end{equation}
so criticality is reached as $g\uparrow g_c$ where the largest real eigenvalue of $\mathbb T$ first reaches 1.
Now, setting $g=g_c$ and working in the critical ensemble, consider the infinite single spine trees with first vertex labelled by $\ell(u_1)=i$. Decomposing these trees at their first vertex $u_1$ into an infinite component and finite components we see that their measures $M_i$ satisfy
\begin{equation}\label{M_evector} M_i=\left(\frac{\partial F_i}{\partial W_j}\right)_c M_j = \mathbb T_{cij} M_j.\end{equation}
We will assume that they are normalised so that the total measure is 1,
\begin{equation} 1=\sum_i M_i.\end{equation}
We see that $\mathbb T$ directly relates the infinite spine trees and the GCE. At criticality it must have at least one eigenvalue which is one and this must be the first real eigenvalue to reach one, otherwise the system would have reached criticality at some smaller value of $g$.
Turning to the local Hausdorff dimension we note from \eqref{Kformula}, \eqref{H3} and \eqref{M_evector} that
$K_i^R$, the expectation value with respect to the measure $\mu_i$ of the number of vertices at distance $R$ from the root of an infinite tree,
is given by
\begin{eqnarray} {\mathbf K}^{R} &=&{\mathbf M} + \sum_{\ell=1}^{R-1} \mathbb T^{\ell-1}\,\Gamma \,\mathbb T^{R-\ell-1}\,{\mathbf W},
\label{Kanswer}\end{eqnarray}
where
\begin{equation} \Gamma_{ik} = M_j\frac{\partial}{\partial W_k}\left(\frac{\partial F_i}{\partial W_j}\right)= M_j\Lambda_{i,jk}. \end{equation}
The implications in general of \eqref{Kanswer} for the Hausdorff dimension depend very much upon the Jordan decomposition of $\mathbb T = S J S^{-1}$ where $J$ is of Jordan Block form $\mathrm{Diag}(J^{<},J^1)$; the block $J^1$ corresponds to $r$ unit eigenvalues and $J^{<}$ to the $N-r$ eigenvalues which are less than 1. If $J^1$ is diagonal then
\begin{equation}\label{Jpower} J^\ell = \mathrm{Diag}(0,\ldots 0,1,\ldots 1) +O(\alpha^\ell)\end{equation}
where $\alpha$ is the largest eigenvalue smaller than 1.
We first consider the case where there is a single unit eigenvalue so $J^1=1$. Introduce the orthonormal basis $(\mbox{e}^i)_j=\delta_{ij}$ so that
\begin{eqnarray} \delta \mathbf{W} &=& S (\mu\, \mbox{e}^N+ \nu^a \,\mbox{e}^a)\,, \qquad
\mathbf{W} = S (A\, \mbox{e}^N+ B^a \,\mbox{e}^a)\end{eqnarray}
where $a, b =1\ldots N-1$. $A\ne 0$ and $B^a$ are constants and up to normalisation the measure vector is
\begin{equation} \mathbf{M}=S\,\mbox{e}^N.\end{equation}
Substituting in \eqref{GCE} we obtain
\begin{eqnarray} (1-J^<)\nu^a \,\mbox{e}^a &=& (A\, \mbox{e}^N+ B^a \,\mbox{e}^a)\delta g+ \frac{1}{2}\mu^2S^{-1}\Gamma S \mbox{e}^N+\mu\nu^aS^{-1}\Gamma S\,\mbox{e}^a\nonumber\\
&&+\frac{1}{2}\nu^a\nu^b S^{-1} \mathbf\Lambda_{jk} (S\,\mbox{e}^a)_j (S\,\mbox{e}^b)_k\,.\label{GCEmcrit}
\end{eqnarray}
Since $1-J^<$ is invertible we see that, provided $ (S^{-1}\Gamma S )_{NN}$ is non-zero,
$\mu\sim (\delta g)^\frac{1}{2}$ and criticality is quadratic. The multi-critical condition is
\begin{equation} (S^{-1}\Gamma S )_{NN}=0.\label{multicrit}\end{equation}
Now returning to \eqref{Kanswer}
\begin{eqnarray} {\mathbf K}^{R}
&=&{\mathbf M} + \sum_{\ell=1}^{R-1} S J^{\ell-1}\,S^{-1}\Gamma \,SJ^{R-\ell-1}S^{-1}\,{\mathbf W}
\end{eqnarray}
whence, using \eqref{Jpower},
\begin{eqnarray} {\mathbf K}^{R}
&=&{\mathbf M} + R \,\mathbf{S} \,(S^{-1}\Gamma \,S)_{NN} (S^{-1}\,{\mathbf W})_N +O(1)
\label{Kresult}\end{eqnarray}
where $ \mathbf{S}_i=S_{iN}$.
We see from \eqref{multicrit} and \eqref{Kresult} that the linear term linear in $R$ automatically vanishes at the multi-critical point where, therefore, $d_h=1$. This is completely standard multi-criticality and the result is identical to that for the single component multi-critical tree. It is straightforward to generalise this analysis to models where $J^1$ is of higher rank but still diagonal and find the same conclusion that $d_h=1$. Note that it is always necessary to compute the actual coefficient of the remaining leading term in a particular model to check that it is positive otherwise the result is meaningless.
The situation is different if $J^1$ is non-diagonal. We will analyse the simplest case where there are two unit eigenvalues and we have the simplest non-diagonal Jordan block
\begin{eqnarray} J^1=\left(\begin{array}{cc} 1&1\\ 0&1\end{array}\right).\end{eqnarray}
We then have that
\begin{equation}\label{JBform} (J^\ell )_{ij}=\delta_{i,N-1}\delta_{j,N-1}+\delta_{i,N}\delta_{j,N}+\ell\delta_{i,N-1}\delta_{j,N} + O(\alpha^\ell).\end{equation}
$\mathbf{M}$ is the ordinary eigenvector with eigenvalue 1,
\begin{equation} \mathbf{M}=S\mbox{e}^{N-1}\end{equation}
but now there is a vector belonging to the second eigenvalue 1
\begin{equation} \boldsymbol \varepsilon=S\mbox{e}^N\end{equation}
with the property
\begin{equation} \mathbb T\, \boldsymbol \varepsilon=\mathbf{M}+\boldsymbol \varepsilon.\end{equation}
Now we have
\begin{eqnarray} \delta \mathbf{W} &=& S (\mu\, \mbox{e}^{N-1}+\lambda\,\mbox{e}^{N}+ \nu^a \,\mbox{e}^a)\, , \quad
\mathbf{W} = S (A\, \mbox{e}^N+C\, \mbox{e}^{N-1}+B^a \,\mbox{e}^a)\,,\end{eqnarray}
where $a, b =1\ldots N-2$. $A\ne 0$ and $B^a$ are constants
and substituting in \eqref{GCE} we find
\begin{eqnarray} (1-J^<)\nu^a \,\mbox{e}^a+\lambda \, \mbox{e}^{N-1} &=& (A\, \mbox{e}^N+C\, \mbox{e}^{N-1}+ B^a \,\mbox{e}^a)\delta g + \nonumber\\
&& +\frac{1}{2} S^{-1}\mathbf\Lambda_{jk}\, (\mu\, \mathbf{M}+ S \lambda\,\mbox{e}^{N}+ \nu^a S\,\mbox{e}^a)_j \times \, \nonumber\\ &&S (\mu\, \mbox{e}^{N-1}+\lambda\,\mbox{e}^{N}\mu\, \mathbf{M}+ \nu^b \mbox{e}^b)_k +{\mathrm {h.o.t.}}\label{GCEmcritJworking}
\end{eqnarray}
and closing with $\mbox{e}^N$,
\begin{eqnarray} 0&=& A\delta g+ \frac{1}{2}\mu^2(S^{-1}\Gamma S)_{N N-1}+\ldots \label{GCEmcritJ}
\end{eqnarray}
showing that
\begin{eqnarray} (S^{-1}\Gamma S)_{N N-1}=0\end{eqnarray}
is a necessary condition for multi-criticality. Note that because of the Jordan block structure $\lambda$ appears linearly on the l.h.s. of \eqref{GCEmcritJworking} so the leading singularity can only be associated with $\mathbf{M}$, and not with $\boldsymbol\varepsilon$.
Using \eqref{Kanswer} and \eqref{JBform} we get
\begin{eqnarray} {\mathbf K}^{R} _i
&=&{\mathbf M}_i + \sum_{\ell=1}^{R-1} S _{ij'}
( \delta_{j',N-1}\delta_{j,N-1}+\delta_{j',N}\delta_{j,N}+(\ell-1)\delta_{j',N-1}\delta_{j,N} )\nonumber\\
&&\,(S^{-1}\Gamma \,S)_{jn}
(\delta_{n,N-1}\delta_{m,N-1}+\delta_{n,N}\delta_{m,N}+(R-\ell-1)\delta_{n,N-1}\delta_{m,N} )
(S^{-1}\,{\mathbf W})_m+O(1)\nonumber\\
&=&{\mathbf M}_i + \frac{1}{6} (R-1)(R-2)(R-3) S_{iN-1} (S^{-1}\Gamma \,S)_{N N-1} (S^{-1}\,{\mathbf W})_N\nonumber\\
&&+\frac{1}{2} (R-1)(R-2) \sum_{L=N-1}^N \left[ S _{iN-1} (S^{-1}\Gamma \,S)_{N L}(S^{-1} \,{\mathbf W})_{L} +
\,S_{iL}(S^{-1}\Gamma \,S)_{L N-1}(S^{-1}\,{\mathbf W})_{N}\right]\nonumber\\
&&+O(R)\label{KmcritJ}
\end{eqnarray}
We see from \eqref{KmcritJ} that the multi-critical condition automatically suppresses the $(R-1)(R-2)(R-3)$ in $\mathbf{K}^R$ but that the quadratic $(R-1)(R-2)$ term survives. Hence $d_h=3$, again provided that the numerical coefficient, which has to be computed in a particular model, is positive.
|
\section{Introduction}
The running of the spectral index has sometimes been used to discriminate
inflationary models.
Recent discovery of B-mode polarization by the BICEP2
collaboration~\cite{Ade:2014xna} would be an important signal of
primordial inflation, since it would indicate the existence of quantum
gravitational radiation, which is a generic prediction of inflationary
cosmological models~\cite{Lyth:1998xn, Olive:1989nu}.\footnote{See also
Ref.~\cite{Mortonson:2014bja}, which shows (in contrast to
Ref.~\cite{Ade:2014xna}) that joint BICEP2
$+$ Planck analysis could favor solutions without gravity waves.}
Although the tensor perturbation is very powerful in discriminating
inflationary models, the BICEP2 result is in some tension with previous
experiments such as WMAP~\cite{Hinshaw:2012aka} and
Planck~\cite{Ade:2013uln}, which claimed upper limits on $r$ (tensor to
scalar ratio).
To resolve the tension, BICEP2 collaboration
suggested~\cite{Ade:2014xna} that the scalar spectral index could run
fast.
However, the signature of the required running $\alpha_s=-0.022\pm 0.010$
($68\%$CL)~\cite{Ade:2014xna} is negative and its absolute value is very
large compared with previous expectation.\footnote{See also recent studies~\cite{recent-st}}
In this paper we show that the constraints from the running spectral
index could be nullified by introducing an extra free scalar field.
Multifield models of inflation may allow inflationary trajectory that is
sensitive to the initial condition.
Recently, it was pointed out that some multifield models make
predictions for the inflationary observables that do not depend strongly
on the specific initial condition~\cite{multi-stat, Easther:2013rva}.
On the other hand, observational data could be used to constrain the
initial field configuration if the model has sensitivity to the
initial condition.
Recent numerical calculation~\cite{Easther:2013rva} clearly shows that
such sensitivity can generate significant running in
double quadratic inflation.\footnote{In this paper we investigate a
possibility of shifting $\alpha_s$ without changing significantly the other
cosmological predictions.
Although Fig.3 of Ref.\cite{Easther:2013rva} is
illuminating, it might be showing that the
model could not give $\alpha_s\sim -0.01$ without shifting $n_s$.}
Our result might indicate that the running spectral index could be a free
parameter that may not be taken too much seriously in this setup.
\section{Running spectral index with a free scalar field}
\subsection{A toy model: Extra component $\dot{\rho}_\chi/\rho_\chi=$const.}
We consider a model in which the curvature
perturbation is sourced by the inflaton $\phi$.
The additional free scalar field is $\chi$, which is dynamical at the
beginning of inflation and may contribute to the adiabatic curvature
perturbation at that moment.
In this section we introduce extra field $\chi$, which will soon
disappear because of $\dot{\rho}_\chi =-6H\rho_\chi<0$.
Note that this toy model has $\ddot{\rho}_\chi>0$, whose contributeion to
the running is usually $\Delta \alpha_s>0$.
Therefore, although the model is simple and very useful for the intuitive
argument, the model is not suitable for generating negative running.
The curvature perturbation at the horizon exit is calculated as
\begin{eqnarray}
\label{zeta-ini}
\zeta&=&-H\frac{\delta
\rho}{\dot{\rho}}\nonumber\\
&=& -H\frac{\delta \rho_\phi+\delta
\rho_\chi}{\dot{\rho}_\phi+\dot{\rho}_\chi}\nonumber\\
&\equiv& r_\phi \zeta_\phi +(1-r_\phi)\zeta_\chi,
\end{eqnarray}
where $\rho_i$ and $P_i$ are the density and the pressure of the component
labeled by $i$.
For the inflaton $\phi$, they are defined as
\begin{eqnarray}
\rho_\phi&\equiv&\frac{1}{2}\left(\dot{\phi}\right)^2
+ V(\phi)\nonumber\\
P_\phi&\equiv&\frac{1}{2}\left(\dot{\phi}\right)^2- V(\phi).
\end{eqnarray}
Here $\rho\equiv \rho_\phi+\rho_\chi$ is defined for the total energy density.
We used component perturbation $\zeta_i$ and the ratio $r_\phi$, which are
defined by
\begin{eqnarray}
\zeta_i&\equiv& -H\frac{\delta \rho_i}{\dot{\rho}_i},\nonumber\\
r_\phi&\equiv& \frac{\dot{\rho}_\phi}{\dot{\rho}_\phi+\dot{\rho}_\chi}.
\end{eqnarray}
If $\chi$ has a flat potential $V(\chi)=0$ and initial velocity
$v_{\chi0}\ne 0$, it will have a decaying velocity
$\dot{\chi}(t)=v_{0}e^{-3Ht}$ and could have a negligible perturbation $\delta \rho_\chi$.
The energy density of $\chi$ obeys $\dot{\rho}_\chi=-3H(1+w)\rho_\chi$
$(w=1)$.
For our calculation we define a ratio between densities: $f\equiv
\rho_\chi/\rho_\phi$, for which inflationary expansion requires $f\ll 1$.
One might argue that $\zeta_\phi$ is not conserved after horizon exit, since the
equation of motion does depend on $\rho_\chi$ via the Hubble parameter
$H^2= \frac{\rho}{3M_p^2}=\frac{\rho_\phi+\rho_\chi}{3M_p^2}$.
The equation of motion is
\begin{eqnarray}
\ddot{\phi}+3H\dot{\phi}+V_{,\phi}=0,
\end{eqnarray}
where the subscript with comma means derivatives with respect to the
field.
Conservation of $\zeta_\phi$ is expected when $P_\phi$ is practically a
unique function of $\rho_\phi$~\cite{New-cons}.
In our scenario, $P_\phi$ might not be an exact unique function of
$\rho_\phi$, however for $f\ll 1$, one can expect that both $P_\phi$ and
$\rho_\phi$ are practically unique functions of $\phi(t)$.
Barring small correction from $f\ll 1$, the curvature perturbation after
inflation will be $\zeta\simeq \zeta_\phi$~\cite{Infcurv}, which has the
spectrum
\begin{eqnarray}
\label{curv-pert-spe}
{\cal P}_{\zeta}&=&
\left(\frac{H}{\dot{\phi}}\right)^2\left(\frac{H}{2\pi}\right)^2
\nonumber\\
&=&\frac{1}{8\pi^2M_p^2}\frac{H^2}{\epsilon_\phi},
\end{eqnarray}
where the slow-roll parameter is defined by
\begin{eqnarray}
\epsilon_\phi &\equiv& \frac{M_p^2}{2}\left(\frac{V_{,\phi}}{\rho}\right)^2.
\end{eqnarray}
Here, $\zeta_\phi$ is defined at the horizon exit.
For later calculation we also define
\begin{eqnarray}
\eta_\phi&\equiv& \frac{V_{,\phi\phi}}{3H^2}.
\end{eqnarray}
We are assuming that the energy density $\rho_\chi$ of the additional
field is a negligible fraction of the total energy density during inflation.
Although $\rho_\chi$ is a small fraction of the total energy density,
$|\dot{\rho}_\chi|\sim |\dot{\rho}_\phi|$ is possible when $\phi$ is
slow-rolling.
(If we make an assumption that our mechanism does not change the original
$n_s-1$, we need to consider $|\dot{\rho}_\chi|\lesssim |\dot{\rho}_\phi|$.
In that case, contribution to $\alpha_s$ will be maximum when
$|\dot{\rho}_\chi|\sim |\dot{\rho}_\phi|$.
We are not fine-tuning the quantities to obtain
$|\dot{\rho}_\chi|\sim |\dot{\rho}_\phi|$, since we have little
difficulty with obtaining the required
$\alpha_s$.)
Moreover, one can easily imagine that $|\ddot{\rho}_\chi|\gg
|\ddot{\rho}_\phi|$ could be possible even if $f\ll 1$.
Our idea is based on that simple speculation.
Since we are introducing a free scalar field $\chi$, $\epsilon_H$ is
splitted into two parts:
\begin{eqnarray}
\epsilon_H&\equiv& -\frac{\dot{H}}{H^2}\nonumber\\
&=&-\frac{\dot{\rho}_\phi}{6H^3 M_p^2}
-\frac{\dot{\rho}_\chi}{6H^3 M_p^2}\nonumber\\
\nonumber\\
&\equiv&\epsilon_{\phi}+\epsilon_{H\chi}.
\end{eqnarray}
Introducing $R\equiv \rho_\chi/3H^2M_p^2$, we find
\begin{eqnarray}
\epsilon_{H\chi} &=& \frac{3(1+w)}{2}R.
\end{eqnarray}
Here $\epsilon_{H\chi}$ should be discriminated from $\epsilon_\chi$,
which will be used in later calculation.
$\epsilon_{H\chi}$ is identical to $\epsilon_{\chi}$ only when $\ddot{\chi}$ is
negligible.
Then, using $d\ln k =Hdt$, we find for the slow-rolling inflaton:
\begin{eqnarray}
\frac{1}{d\ln k}\epsilon_\phi
&=&4\epsilon_\phi\epsilon_H-2\eta_\phi\epsilon_\phi.
\end{eqnarray}
In the first term, $\epsilon_H$ appears instead of $\epsilon_\phi$, since
our definition of $\epsilon_\phi$ uses $\rho$ instead of $V$.
For our calculation we also evaluate
\begin{eqnarray}
\frac{1}{d\ln k}\epsilon_{H\chi}&=&
\frac{3(1+w)}{2H}\dot{R}
\nonumber\\
&=&\frac{3(1+w)}{2H}\left[R\frac{\dot{\rho}_\chi}{\rho_\chi}-2R\frac{\dot{H}}{H}\right]
\nonumber\\
&=&-3(1+w)\epsilon_{H\chi}+2\epsilon_{H\chi}\epsilon_H.
\end{eqnarray}
The spectral index of the curvature perturbation is
\begin{eqnarray}
\label{eq-index}
n_s-1&\equiv&\frac{d\ln {\cal P}_\zeta}{d\ln k}\nonumber\\
&=&-6\epsilon_H+2\eta_\phi.
\end{eqnarray}
Since $\epsilon_H=\epsilon_\phi+\epsilon_{H\chi}$, the shift caused by
the additional field $\chi$ is
\begin{eqnarray}
\Delta (n_s-1)&=&-6\epsilon_{H\chi}.
\end{eqnarray}
Since we are considering a conservative scenario in which additional field $\chi$ does not
change the original index, we need $6\epsilon_{H\chi}\ll 0.04$.
We also have
\begin{eqnarray}
\frac{1}{d\ln
k}\eta_\phi&=&\frac{1}{Hdt}\left[\frac{V_{,\phi\phi}}{3H^2}\right]\nonumber\\
&=&\frac{1}{H}\left(\frac{V_{,\phi\phi\phi}\dot{\phi}}{3H^2}\right)+2
\left(\frac{V_{,\phi\phi}}{3H^2}\right)\left(-\frac{\dot{H}}{H^2}\right)\nonumber\\
&=&-\left(\frac{V_{,\phi\phi\phi}V_{,\phi}}{9H^4}\right)+2
\left(\frac{V_{,\phi\phi}}{3H^2}\right)\left(-\frac{\dot{H}}{H^2}\right)\nonumber\\
&=&-\xi_\phi +2\eta_\phi\epsilon_H,
\end{eqnarray}
where we defined
\begin{eqnarray}
\xi_{\phi}&\equiv& \frac{V_{,\phi} V_{,\phi\phi\phi}}{9H^4}.
\end{eqnarray}
Then, the running of the spectral index is
\begin{eqnarray}
\label{alpha-eq-kin}
\alpha_s&\equiv&\frac{d n_s}{d\ln k}
\nonumber\\
&=& -6\left[4\epsilon_\phi\epsilon_H-2\eta_\phi\epsilon_\phi\right]\nonumber\\
&&+2\left[2\eta_\phi\epsilon_H -\xi_\phi\right]\nonumber\\
&&-6\left[-3(1+w)\epsilon_{H\chi}+2\epsilon_{H\chi}\epsilon_H\right]
\nonumber\\
&\simeq& \left[-24\epsilon_\phi\epsilon_H+16\eta_\phi\epsilon_\phi -2\xi_\phi\right]
\nonumber\\
&&+4\eta_\phi\epsilon_{H\chi}+18(1+w)\epsilon_{H\chi}-12\epsilon_{H\chi}\epsilon_H,
\end{eqnarray}
where we can see correction $\Delta \alpha_s\sim 18(1+w)\epsilon_{H\chi}\lesssim
0.12(1+w)$, but the sign of this term is positive.
We will see that the correction can be negative when $\chi$ is moving on
a hilltop potential.
For our scenario, we are considering an inflationary model in which
final $\zeta$ is determined by $\phi$, but an extra
component may change scale dependence of the spectrum.
Multi-field models of inflation have been studied by many
authors~\cite{Polarski:1992dq, Bassett:2005xm, Vernizzi:2006ve}.
From those studies one can see that an additional field, which could be
dynamical at the horizon exit but will soon stop or
slow down, may not change final $\zeta$ at least at the first order.
The usual calculation uses adiabatic and entropy perturbations instead of
component perturbations, and considers conversion between them.
Although in this paper we will consider
calculation that is usually considered for the curvaton
mechanism~\cite{Infcurv}, one will reach the same result using the calculation
considered in Ref.\cite{Polarski:1992dq, Bassett:2005xm,
Vernizzi:2006ve}.
A similar idea has been studied when there is no extra field but there
is a deviation from the
slow-roll velocity~\cite{Damour:1997cb, Liddle:1998pz,
Taruya:1998cz, Yokoyama:1998rw, Seto:1999jc, Takamizu:2010je}.
In those studies it has been found that the curvature perturbation could
not be affected by such additional degrees of freedom if they are
disappearing during inflation.
Also in Ref.~\cite{matsuda-elliptic, Matsuda:2008fk, Matsuda:2012kk},
it has been pointed out that such field might shift the spectral index and
its running.
Viewing those studies, we think that the idea of changing the scale
dependence of the curvature perturbation using an additional dynamical
field could have been known partly, although we could not find explicit
calculation that can be used for solving the tension between BICEP2 and
other experiments.
\subsection{slow-roll $\chi$}
In the above calculation we considered an additional component that
obeys $\dot{\rho}_\chi=-3H(1+w)\rho_\chi$.
Let us see what happens if $\chi$ is a moderately slow-rolling field
(but it will soon reach its minimum or will soon be negligible).
Using conventional slow-roll parameters for both $\phi$ and $\chi$,
we find the spectral index
\begin{eqnarray}
\label{eq-index2}
n_s-1&=&-6\epsilon_H+2\eta_\phi,
\end{eqnarray}
where $\epsilon_H\equiv \epsilon_\phi+\epsilon_\chi$.
Here we define $\epsilon_\chi\equiv
\frac{M_p^2}{2}\frac{V_{,\chi}}{\rho}$, which is equivalent to
$\epsilon_{H\chi}$ when $\ddot{\chi}$ is negligible.
The running of the spectral index is
\begin{eqnarray}
\label{alpha-eq}
\alpha_s
&=& -6\left[4\epsilon_\phi\epsilon_H-2\eta_\phi\epsilon_\phi\right]\nonumber\\
&&+2\left[2\eta_\phi\epsilon_H -\xi_\phi\right]\nonumber\\
&&-6\left[4\epsilon_\chi\epsilon_H-2\eta_\chi\epsilon_\chi\right]
\nonumber\\
&\simeq& \left[-24\epsilon_\phi\epsilon_H+16\eta_\phi\epsilon_\phi -2\xi_\phi\right]
\nonumber\\
&&+4\eta_\phi\epsilon_\chi-24\epsilon_\chi\epsilon_H
+12\eta_\chi\epsilon_\chi,
\end{eqnarray}
where the major correction appears in $|12\eta_\chi\epsilon_\chi|.$
If we assume $|\Delta (n_s-1)|=6\epsilon_\chi<|n_s-1|$, we will
have $|\Delta \alpha_s|<|2\eta_\chi
(n_s-1)|\sim 0.08|\eta_\chi|$.
Since $\eta_\chi$ does not appear in the spectral index, there is no
bound for $|\eta_\chi|$.
Therefore, one can take $|\eta_\chi|$ as large as $0.125$, which gives
correction as large as $|\Delta \alpha_s|\sim 0.01$.
Since we are considering an extra dynamical field that will soon
disappear,
$\eta_\chi\sim 0.125$ is a conceivable choice for the model.
Since $\eta_\chi<0$ is possible for a hilltop potential, one will be able to find
a correction $\Delta \alpha_s \sim -0.01$.
\subsection{fast-roll $\chi$}
Alternatively, one can consider a hilltop potential $V(\chi)=
-\frac{1}{2}m_\chi^2 \chi^2$ for a fast-rolling field $\chi$ and will
find~\cite{Dimopoulos:2003ce}
\begin{eqnarray}
\chi(t)&=& \chi(t_0)e^{Kt},
\end{eqnarray}
where
\begin{eqnarray}
K&\equiv&
\frac{3}{2}H\left[1-\sqrt{1-\frac{4}{9}\left(\frac{m_\chi^2}{H^2}\right)}
\right].
\end{eqnarray}
Conventional fast-rolling condition is ``$\epsilon_\chi\ll
1$ but $|\eta_\chi|\sim {\cal O}(1)$''.
Then the slow-roll parameter can be replaced as
\begin{eqnarray}
\epsilon_{H\chi}&=&-\frac{\dot{\rho}_\chi}{6H^3M_p^2}\nonumber\\
&=&\frac{(-K^3+K m_\chi^2)\chi^2 }{6H^3M_p^2}\nonumber\\
&=&\left(-\frac{K^2}{m_\chi^2}+1\right)\frac{K}{H}R,
\end{eqnarray}
where contribution from the kinetic term has not been neglected since
$\ddot{\chi}$ could not be negligible for the fast-rolling field.
Then we find
\begin{eqnarray}
\frac{1}{d\ln k}\epsilon_{H\chi}&=&
-\frac{\ddot{\rho}_\chi}{6H^4M_p^2}
+3\epsilon_{H\chi}\epsilon_H
\nonumber\\
&=&\frac{2K}{H}\epsilon_{H\chi}
+3\epsilon_{H\chi}\epsilon_H.
\end{eqnarray}
The spectral index is
\begin{eqnarray}
n_s-1&=&-6\epsilon_\phi+2\eta_\phi
-6\epsilon_{H\chi},
\end{eqnarray}
which gives $6\epsilon_{H\chi}<0.04$ for our scenario.
The running of the spectral index is
\begin{eqnarray}
\alpha_s&=&-6[4\epsilon_\phi\epsilon_H-2\eta_\phi\epsilon_\phi]
+2[2\epsilon_H\eta_\phi-\xi_\phi]\nonumber\\
&&-6\left[\frac{2K}{H}\epsilon_{H\chi}
+3\epsilon_{H\chi}\epsilon_H\right]\nonumber\\
&=& \left[-24\epsilon_\phi\epsilon_H+16\eta_\phi\epsilon_\phi -2\xi_\phi\right]
+4\epsilon_{H\chi}\eta_\phi\nonumber\\
&&-6\left[\frac{2K}{H}\epsilon_{H\chi}
+3\epsilon_{H\chi}\epsilon_H\right].
\end{eqnarray}
The correction appears in $-12\epsilon_{H\chi}K/H$, where $\Delta
\alpha_s$ is negative for the hilltop potential.
If one takes $K/H\sim 1$ (fast-roll), one will find
$\Delta \alpha_s\sim -0.01$ for $\epsilon_{H\chi}\sim 8\times 10^{-4}$.
\subsection{The Curvaton}
For the conventional curvaton mechanism, one will
find~\cite{Kobayashi:2012ba}
\begin{eqnarray}
\alpha_s&\simeq&2\frac{\ddot{H}}{H^3}-4\epsilon_H^2
+4\epsilon_H\eta_\sigma
-2\xi_\sigma,
\end{eqnarray}
where $\sigma$ is the curvaton.
If we introduce an extra light field $\chi$, significant correction may
appear from the first term, which gives for a fast-rolling $\chi$:
\begin{eqnarray}
2\frac{\ddot{H}}{H^3}&=& \frac{\ddot{\rho}_\chi}{3H^4M_p^2}+...\nonumber\\
&\sim&2\left(1-\frac{K^2}{m_\chi^2}\right)\frac{K^2}{H^2}R.
\end{eqnarray}
Therefore, one can easily find similar correction in various other models including
the curvaton.
\subsection{A model}
In this section we show a specific scenario in which $\alpha_s$ can be
shifted by an additional field $\chi$.
At the beginning of inflation, the additional field has a hilltop potential
\begin{eqnarray}
V(\chi)=-\frac{1}{2}m_\chi^2\chi^2,
\end{eqnarray}
and $\chi$ will soon reach the minimum at $\chi_0$.
Here we assume $\chi_0\sim M_p$.
The duration is bounded by $e^{K\Delta t}=\frac{\chi_\mathrm{end}}{\chi_\mathrm{ini}}<\frac{M_p}{H}\simeq
2.4\times 10^{4}$~\cite{Ade:2014xna},
where
$\chi_\mathrm{ini} > \delta \chi \simeq H/2\pi$ and
$\chi_\mathrm{end}<\chi_0\sim M_p$ has been considered.
We find
$\Delta N\lesssim 10\times\left(\frac{H}{K}\right)$.
The additional field $\chi$ must reach its minimum before the end of inflation.
For a moderately slow-rolling $\chi$, $\epsilon_{\chi}$ is determined by
$m_\chi$ and $H$.
Then $|\Delta (n_s-1)|<0.04$ gives for the quadratic potential:
\begin{eqnarray}
6\epsilon_\chi&=&6\times \frac{M_p^2}{2} \left(\frac{-m_\chi^2
\chi}{\rho}\right)^2_*
=6\times
\left(\frac{-m_\chi^2}{3H^2}\right)_*\left(\frac{-\frac{1}{2}m^2_{\chi}\chi^2}{\rho}\right)_*
\nonumber\\
&=&6\eta_\chi R_*<0.04.
\end{eqnarray}
Here the quantities with star is defined
when the perturbation leaves horizon.
We consider $\eta_\chi\sim 0.6$ and $|R_*|\sim 0.01$ on the scale where
significant $\alpha_s$ could be observed.
The condition $|R_*|\sim 0.01$ is conceivable since it is required at the
very early stage of inflation.
Then, we can estimate the correction as
\begin{eqnarray}
|\Delta \alpha_s|&\simeq & |12\epsilon_\chi\eta_\chi|=|12\eta_\chi^2
R|\sim 0.04.
\end{eqnarray}
In the same way, the running of $\alpha_s$ will be shifted by a similar
order, which is consistent with Planck
experiment.\footnote{See Fig.3 of Ref.~\cite{Ade:2013uln}.}
On the other hand, one might suspect that higher
runnings may ovecome the correction coming form $\alpha_s$.
The expansion will be~\cite{Kohri:2013mxa}
\begin{eqnarray}
{\cal P}_\zeta(k)&=&{\cal P}_\zeta(k_0)\exp\left[
(n_s-1)\ln\left(\frac{k}{k_0}\right)
+\frac{\alpha_s}{2}\ln^2\left(\frac{k}{k_0}\right)\right.\nonumber\\
&&\left.+\frac{\beta_s}{3!}\ln^3\left(\frac{k}{k_0}\right)
+...\right]\nonumber\\
&=&
{\cal P}_\zeta(k_0)
\left(\frac{k}{k_0}\right)^{
(n_s-1)
+\frac{\alpha_s}{2}\ln\left(\frac{k}{k_0}\right)
+\frac{\beta_s}{3!}\ln2\left(\frac{k}{k_0}\right)
+...}.
\end{eqnarray}
Note that there will be no problem in the expansion if one takes
$\ln(k/k_0)=\ln(0.05/0.0027)\simeq 2.92$ and
$\beta_s < \alpha_s$, where $\beta_s$ is the running of $\alpha_s$.
Our model suggests that $\alpha_s$ and $\beta_s$ could be a similar
order.
However, normally $\beta_s$ does not exceed $\alpha_s$, and the higher
runnings does not
spoil the scenario
of nullifying the tension between BICEP2 and
the Planck experiments.
We are anticipating that there could be some significant sign of new
physics in the higher runnings, however at this moment higher
runnings are not so much constrained.
Such a large running (and a running of running) will be checked by
the future B-mode polarization and the 21cm line
observations.~\cite{Kohri:2013mxa}.
One may use this to distinguish the source of the scale
dependence if a signature might be observed in future
experiments.
\section{Conclusion and discussion}
\label{conclusion}
Normally in single-field inflation models, we have only one
scalar field that is responsible for both the inflationary expansion and
the production of the curvature perturbation.
In reality, however, there could be many other scalar fields {\it besides} the
inflaton field, which could be dynamical at least at the beginning of
the inflationary stage.
Although the non-inflaton components may disappear (or becomes static)
before the end of inflation, those additional fields could change the
scale dependence of the spectrum.
If one wants to remove those extra fields before the onset of $N_e\sim
60$ inflation, one can assume extra e-foldings prior to $N_e\sim60$, so
that such fast-rolling (or modestly fast-rolling) fields are already
negligible from the beginning.
In that case, remaining fields will not have significant
contribution to $\alpha_s$.
On the other hand, in the light of BICEP2 and the Lyth bound, such extra e-foldings would
require more excursion of the inflaton field, which could be a problem
if one takes the trans-Planckian problem seriously.
We thus expect that such extra dynamical degrees of freedom can exist
naturally in modern inflationary scenarios.
As long as we know, the idea has not been discussed yet to
solve the tension between BICEP2 and other experiments.
In this paper we considered a free scalar field and calculated its
contribution to $n_s$ and $\alpha_s$.
We assumed conservatively that spectral index is not much altered by the
additional field.
Using a simple model, we showed that $O(0.01)$ ambiguity can be found in
the running spectral index if an extra field could be dynamical when the
perturbation leaves horizon.
In the same way, the running of $\alpha_s$ can be shifted by a similar
order, which could be used to distinguish the source of the scale
dependence.
At the same time, the scale-dependence of the tensor mode could be
affected, which has already been discussed in
Ref.\cite{Gerbino:2014eqa, Gong:2014qga, Kobayashi:2013awa}.
In our scenario, we find $r\sim 16 \epsilon_\phi$, which does not
include $\epsilon_\chi$.
This means that $r$ is not affected by $\epsilon_\chi\ne 0$.
Similarly, the running of $r$ is determined by $\dot{\epsilon}_\phi$,
which does not introduce significant running at the first
order.
Significant running may appear from the next order of the running.\footnote{The second order (i.e, the running of of running of $r$)
can be large since it will include $\ddot{\epsilon}_\phi$.}
On the other hand, in our model $\frac{d n_t}{d\ln k}\propto \dot{\epsilon}_H$
can be large.
Since the scale-dependence of the tensor power spectrum is signicant
in our case, while such a scale dependence is usually not assumed,
there could be concern
that it might bias the estimation of parameters like $r$ and
$\alpha_s$ from the WMAP/Planck data. Those quantities could be
negligible at the end, but the resolution of the tension between
BICEP2 and the WMAP/Planck could be affected.
In that sense we need further investigation to obtain more accurate
estimation of the parameters when the tensor mode is
running.\footnote{We thank the reviewer of the journal for these points.}
\section*{Acknowledgement}
K.K. is supported in part by Grant-in- Aid for Scientific
research from the Ministry of Education, Science, Sports, and
Culture (MEXT), Japan, Nos. 21111006, 23540327, 26105520 (K.K.). The
work of K.K. is also supported by the Center for the Promotion of
Integrated Science (CPIS) of Sokendai (1HB5804100).
|
\section{Introduction}
Observational evidences for multiple stellar populations in globular
clusters (GCs) challenge the traditional view of GCs hosting simple
stellar populations, which are formed in one generation from a cloud
of uniform composition (single-generation cluster). One important
evidence is the presence of star-to-star abundance variations. The
most famous example is the Na-O anticorrelation that oxygen-depleted
stars have higher sodium abundances, which is exhibited among the
main sequence (MS) turnoff (TO) stars, sub-giant branch (SGB) stars
and red-giant branch (RGB) stars in some GCs
\citep{Carretta2004,Gratton2004}. Another important evidence is that
some GCs show multiple MSs \citep{Piotto2007}, multiple SGBs
\citep{Milone2008,Marino2009}, or anomalously wide RGB
\citep{Yong2008,Lee2009} in the Hertzsprung-Russell (HR) diagram.
Several scenarios have been proposed to explain the formation of
multiple stellar populations \citep{Gratton2012}, and they can be
classified into two contrasting hypotheses: \emph{Primordial}
Scenario vs. \emph{Evolutionary} Scenario
\citep{Kraft1994,Carretta2004}. In the \emph{Primordial} Scenario,
multiple stellar populations in GCs are formed in the different
materials that are primordial chemical inhomogeneities, for example,
a merger of two separate GCs \citep{Lee1999,Mackey2007} or the
self-pollution of the intra-cluster gas occurring in the early
evolution of clusters
\citep{D'Antona2002,Decressin2007,deMink2009a}. However, a merger
event would not be expected to occur in the halo of the Milky Way
due to the very large relative velocities of GCs
\citep{Gratton2012}. For the the self-enrichment of GCs, it is
difficult to explain the high fraction of the second-generation
formed from the gas polluted by the first-generation stars
\citep{Bastian2009}, and the nature of possible polluters remains
unclear \citep{Sills2010,Gratton2012}.
In the \emph{Evolutionary} Scenario, initial GCs are believed to be
single-generation clusters, and multiple stellar populations in GCs
are attributed to the evolution of stars affected by some physical
processes, such as the first dredge-up, Sweigart-Mengel Mixing
\citep{Sweigart1979}, primordial rotation \citep{Bastian2009}.
However, it is necessary to explain why only some of stars in GCs
are affected by these physical processes. Moreover,
\citet{Decressin2007} suggested that the MS TO stars in GCs are not
hot enough for producing the observed Na-O anti-correlation, because
the central temperature (about 25 $\times$ $10^6$\,K) of a
0.85\,M$_{\rm \odot}$ TO stars is higher than the required
temperature of CNO cycle (about 20 $\times$ $10^6$\,K), but is lower
than that of the production of $^{23}$Na from proton-capture on
$^{20}$Ne (about 35 $\times$ $10^6$\,K).
Binary interactions have been used to explain the formation of
multiple stellar populations as a kind of \emph{Primordial}
Scenario. Chemically enriched material ejected from massive binaries
have been proposed to form a second population of stars
\citep{deMink2009a}, or to be accreted by low-mass pre-main-sequence
stars \citep{Bastian2013}. However, binary interactions are not
considered as a kind of \emph{Evolutionary} Scenario. They are not
investigated to produce the abundance anomalous stars
\emph{directly}, although they have been proposed as an explanation
for extended MS in the HR diagrams of intermediate age clusters
\citep{Yang2011,Li2012}.
Binary systems can produce stars more massive than normal single
stars in the same evolutionary stage by binary interactions, e.g.
the merged stars or the accretor stars. These stars are not hot
enough for the production of $^{23}$Na from proton-capture on
$^{20}$Ne, but they are hot enough for proton-captures on $^{16}$O
and $^{22}$Ne, which respectively lead to the destruction of
$^{16}$O and to the production of $^{23}$Na at temperatures $>$20
$\times$ $10^6$\,K \citep{Prantzos2006}. Furthermore, these stars
produced by binary interactions will be rapid rotating because
binary interactions can convert orbital angular momentum into their
spin angular momentum \citep{deMink2013}. Rotationally induced
mixing can bring part of the nuclear-processed products from the
core to the surface \citep{Decressin2007}, and stellar rotation was
used to explain the observed nitrogen enrichment found in several
massive MS stars \citep{Brott2009}. Therefore, these rapidly
rotating stars reproduce the properties of the abundance anomalous
stars observed in GCs (e.g. the Na-O anticorrelation), and binary
interactions will be a possible scenario for the formation of
multiple stellar populations in GCs as a kind of \emph{Evolutionary}
Scenario.
In this paper, we adopt a simple model for this scenario to
investigate the formation of multiple stellar populations. In
Section 2, we explore the formation of rapidly rotating stars from
binary interactions by performing binary evolution calculations, and
calculated the subsequent evolution of these rapidly rotating stars
to obtain their parameters (e.g. effective temperatures,
luminosities and surface abundances). Then, we estimate the
distributions of the binary population by performing the binary
population synthesis study. In Section 3, we give discussion and
conclusions.
\section{THE MODEL AND SIMULATION}
In our scenario, the anomalous stars observed in GCs are the merged
stars and the accretor stars produced by binary interactions, and
these binary products are rapidly rotating and more massive than the
normal stars. These binary products have different effective
temperatures, luminosities and surface abundances from single stars
in the same evolutionary stage due to their own evolution. Thus, the
stellar population with binaries (the binary population) can
reproduce the observations of multiple stellar populations, because
the binary population includes two populations: (1) normal single
stars with normal abundance (initial abundance of GCs); (2) the
merged stars and the accretor stars (produced by binary
interactions) that have abnormal surface abundances by their own
evolution.
\subsection{Binary evolution calculations}
We considered three evolutionary channels for the formation of
rapidly rotating stars from binary interactions: (1) a binary
evolves into contact and then merges into a rapidly rotating star
(contact merger channel for the merged stars)
\citep{Jiang2012,deMink2013}; (2) a binary evolves into Roche lobe
overflow (RLOF) and experiences unstable mass transfer while both
components are still MS stars, and then merges into a rapidly
rotating star (unstable RLOF merger channel for the merged stars)
\citep{Jiang2012}; (3) a binary evolves into RLOF and experiences
stable mass transfer, and the secondary obtains the mass and angular
momentum transferred from the primary, and then becomes a rapidly
rotating star (stable RLOF channel for the accretor stars)
\citep{deMink2013}.
In these channels, a binary needs to evolve into contact phase or
unstable RLOF phase while both components are still MS stars, or
evolve into stable RLOF phase while the secondary is still a MS star
\citep{Jiang2012,deMink2013}. This is because for MS stars, the
binding energy of the envelope might be large enough to form rapidly
rotating stars. To determine whether the binary evolves into these
phase, it is necessary to perform detailed binary evolution
calculations. We use Eggleton's stellar evolution code
\citep{Eggleton1971,Eggleton1972,Eggleton1973} that has been updated
with the latest input physics during the last four decades
\citep{Han1994,Pols1995,Pols1998,Nelson2001,Eggleton2002}. In our
calculations, we used the metallicity $Z = 0.001$ (corresponding to
typical GCs) and considered the nonconservative effects, includes a
model of dynamo-driven mass loss, magnetic braking and tidal
friction for the evolution of stars with cool convective envelopes
\citep{Eggleton2002}.
Altogether, we have calculated the evolution of 5200 binary systems,
thus obtained a large, dense model grid that covers the following
ranges of initial primary mass ($M_{10}$), initial mass ratio ($q_0
= M_{20}/M_{10}$) and initial orbital period ($P_0$):
log\,$M_{10}=-0.3, -0.25, . . . , 0.2$;
log\,$q_0=$log\,$(M_{20}/M_{10})=0.05, 0.10, . . . , 0.5$; and
log\,$(P_0/P_{\rm ZAMS})= 0.05, 0.10, . . . $, where $P_{\rm ZAMS}$
is the orbital period at which the initially more massive component
would just fill its Roche lobe on the zero-age MS. These ranges of
initial parameters cover the systems that can form rapidly rotating
stars in old stellar population by Case A binary evolution.
According to these binary models, the parameters of rapidly rotating
stars are established when the binary evolves into contact, unstable
RLOF, or the end of stable RLOF. For simplicity, in the contact
merger channel and the unstable RLOF merger channel, we assume that
there are 10\% mass loss during the merger phase
\citep{Jiang2012,deMink2013} and the binary products rotate at 80\%
of break-up velocity. For the stable RLOF channel, we calculate the
angular momentum obtained by the secondary, and then calculate its
rotational velocity. We only consider the secondaries that can get
enough angular momentum to rotate at 80\% of break-up velocity.
\subsection{Rotation evolution calculations}
To calculate the subsequent evolution of rapidly rotating stars
produced by binary interactions, we have used the Modules for
Stellar Experiments in Astrophysics code (MESA version 4631)
\citep{Paxton2011}. This code implements the effects of rotation in
the transport of angular momentum and chemical mixing, and it has
been used to study the effects of rotational mixing
\citep{Chatzopoulos2012a,Chatzopoulos2012b}. We do not consider the
effect of the helium-rich core of the original components on the
subsequent evolution of rapidly rotating stars, and we assume that
these rapidly rotating stars have abundances similar to zero age MS
stars. In addition, we consider that the effect of rotation on
low-mass stars with convective envelopes might be weaker due to
magnetic braking \citep{Hurley2000,Bastian2009}. We calculate the
stellar evolutionary models with $M = 0.6, 0.7 ... 1.6$\,M$_{\rm
\odot}$ for $Z = 0.001$ that cover rapidly rotating stars produced
by binary interactions in old stellar populations. For simplicity,
these models run for two different degrees of rotation: non-rotating
and 80\% of break-up velocity, which are used to investigate the
evolution of surface abundances of single stars and rapidly rotating
stars produced by binary interactions. All evolutionary tracks have
been computed up to the point of core helium ignition.
To illustrate the evolutionary tracks of these stars, we show four
evolution examples of surface mass fractions (sodium versus oxygen
abundances) in Fig. 1. During the evolution of rapidly rotating star
with $M=1.3$\,M$_{\rm \odot}$ (dotted line), the surface mass
fraction of sodium ($X_{\rm Na23}$) increases from $1.22\times
10^{-6}$ to $5.95\times 10^{-6}$ , and that of oxygen ($X_{\rm
O16}$) decreases from $4.67\times 10^{-4}$ to $9.54\times 10^{-5}$.
This is because the destruction of O and the production of $^{23}$Na
in the central region result from proton-captures on $^{16}$O and
$^{22}$Ne \citep{Langer1993,Prantzos2006,Decressin2007}, and they
are brought to the surface by rotational mixing. In other examples,
rapidly rotating stars with $M=1.0, 1.1, 1.2$\,M$_{\rm \odot}$
evolve in a similar way as in the previous example, although the
variation ranges of surface abundances decrease with decreasing
mass. Hence, these rapidly rotating stars can be oxygen depletion
and sodium enhancement. For single stars with no rotation (open
circles in Fig. 1), the surface mass fractions of sodium and oxygen
do not show significant variations.
\subsection{Binary population synthesis}
In the binary population synthesis study, we have performed a Monte
Carlo simulation. We adopt the following input for the simulation
(see Han et al. 1995). (1) The initial mass function (IMF) of
\citet{Miller1979} is adopted. (2) The mass-ratio distribution is
taken to be constant. (3) We take the distribution of separations to
be constant in log\,$a$ for wide binaries, where $a$ is the orbital
separation. Our adopted distribution implies that $\sim$50 per cent
of stellar systems are binary systems with orbital periods less than
100\,yr. We follow the evolution of a million sample binaries
according to grids of binary models and the evolution models of
rapidly rotating stars described above.
The simulation gives the distributions of many properties of the
binary population, e.g. the masses, the effective temperatures, the
luminosities, the surface abundances, etc. Fig. 2 are selected
distributions of the binary population at 10 Gyr that may be helpful
for understanding the scenario of binary interactions for the
formation of multiple stellar populations. Fig. 2(a) shows the
distribution of single stars and binary products in the Na-O plane.
Single stars (open circles) are concentrated at the bottom right
corner and do not show significant variations. However, binary
products have large star-to-star variations. The distribution of
binary products first extends upward to high sodium abundance region
from the original abundance location, and then turns left to the
oxygen depletion region. Thus, the binary population with binary
interactions can show a Na-O anticorrelation.
To investigate the distributions of binary products with different
abundances in HR diagram, we classify them into four groups based on
the surface mass fraction of Na and O, which are denoted as pluses
with different colours. The distribution of the binary population
with binary interactions in HR diagram is shown in Fig. 2(b). It is
obvious that this population has at least two MSs, three TOs, three
SGBs and extended RGB, which are produced by single stars (open
circles) and binary products (pluses). More importantly, these
sequences have a large spread in sodium and oxygen abundances. This
is consistent with the fact that the Na-O anticorrelation is
observed among MS TO stars, SGB stars and RGB stars in some GCs
\citep{Carretta2004,Gratton2004}. Observed red giants in 17 GCs
given by \citet{Carretta2009} are shown in Fig. 2(c), and they have
a distribution similar to that of our simulated binary populations.
If we assume that the binary populations (black line and blue lines)
with different initial abundances, such as in different GCs, have
the same variation ranges of surface abundances, then we would find
that these binary populations in the simulation are in good
agreement with the observed stars in 17 GCs as shown in Fig. 2(c).
We noted that single stars and binary products in the binary
population have different mass distributions as shown in Fig. 2(d),
and binary products are more massive than single stars. For the
binary population with log $L/L_{\rm \odot}
> 0$, the mass range of single stars is from 0.75 to 0.9\,M$_{\rm \odot}$,
while the mass range of binary products is from 0.8\,M$_{\rm \odot}$
to 1.35\,M$_{\rm \odot}$, although these single stars and binary
products have the same range of evolutionary stages. The maximum
mass of binary products is larger than that of single stars by
$\sim0.45$\,M$_{\rm \odot}$. We do not compare stars with log
$L/L_{\rm \odot} < 0$ because the sample of very low-mass binary
products is not complete in our calculations. Although the effect of
rotation on low-mass stars might be weaker due to magnetic braking
\citep{Hurley2000,Bastian2009}, massive binary products could have
significantly different surface abundances from single stars. Binary
products are more massive than single stars in the same evolutionary
stage, and they rotate rapidly, i.e. close to break-up velocity. The
destruction of $^{16}$O and the production of $^{23}$Na in the
central region result from proton-captures on $^{16}$O and
$^{22}$Ne, and the products of nuclear burning are brought to the
surface by rotational mixing. This can explain why the abundance
variations take place in these stars, and not in single stars at the
same evolutionary stage.
\section{Discussion and conclusions}
In this paper, we presented binary interactions as a possible
scenario for the formation of multiple stellar populations in GCs.
In this scenario, binary interactions can convert orbital angular
momentum into spin angular momentum and produce rapidly rotating
stars. These stars might have different properties from single
stars, such as surface abundances, temperatures and luminosities,
because they experienced binary interactions and their evolution is
affected by rotationally induced mixing. The existence of binary
products and single stars results in multiple stellar populations in
GCs, although initial GCs are single-generation clusters. In our
simple model, we carried out binary evolutionary calculations and
constructed the evolutionary models of rapidly rotating stars formed
by binary interactions. By performing a Monte Carlo simulation, we
obtained the distributions of the binary population and investigated
the formation of multiple stellar populations from binary
interactions.
Stellar rotation have been shown that can explain the broad MS and
spread TO observed in intermediate age stellar clusters
\citep{Bastian2009, Li2012}. However, \citet{Bastian2009} suggested
that this is unlikely to cause multiple stellar populations observed
in old GCs, because the stars in GCs are low mass and are not
expected to be rapidly rotating stars. We show that old binary
population could show multiple sequences in HR diagram as shown in
Fig. 2(b), because binary interactions could produce rapidly
rotating stars. Binary interactions, including mass transfer
\citep{Pols1991,deMink2013} or merger \citep{Tylenda2011,Jiang2013},
have been used to investigate the formation of rapidly rotating
stars. Moreover, observed evidences have shown that binary
interactions are very common in GCs. Many contact binaries have been
observed in GCs \citep{Rucinski2000}, and these binaries are
believed to merge into a rapidly rotating star (contact merger
channel). \citet{Mathieu2009} show that many of blue stragglers are
rotating faster than normal MS stars , which might be formed by mass
transfer \citep{Geller2011}. Therefore, binary interactions could
explain multiple sequences in HR diagram, which is important
evidence of multiple stellar populations.
Other important evidence of multiple stellar populations in GCs is
the presence of star-to-star abundance variations, and the most
famous example is the Na-O anticorrelation. We show that old binary
population can reproduce a Na-O anticorrelation as a result of the
formation of rapidly rotating stars from binary interactions. If the
binary populations with different initial abundances, such as in
different GCs, are assumed to have the same variation ranges of
surface abundances, we found that these binary populations in the
simulation are in good agreement with the observed stars in 17 GCs
given by \citet{Carretta2009}. In our scenario, this anticorrelation
could be found in TO stars, SGB stars and RGB stars, which is in
agreement with the observation that show the Na-O anticorrelation in
these evolutionary stages \citep{Carretta2004,Gratton2004}.
Moreover, our model predicts that these abundance anomalous stars
should be more massive than MS TO stars, and they should be blue
stragglers or blue straggler progeny. Therefore, binary interactions
could explain the existence of the Na-O anticorrelation in GCs. This
suggests that binary interactions may be a possible scenario for the
formation of multiple stellar populations, and the traditional view
of single-generation cluster could explain the photometric and
spectroscopic observations of multiple stellar populations by
considering the effects of binary interactions.
Observed multiple stellar populations show the cluster-to-cluster
variations \citep[e.g.][]{Carretta2009}, and these variations also
need to be explained \citep{Sills2010}. This cluster-to-cluster
variations might be due to the dependence of binary interactions on
the cluster properties, such as age, initial abundance, initial
fraction of binaries and initial orbital separation distribution.
These parameters determine the nature of binary interactions and the
subsequent evolution of binary products. When the initial fraction
of binaries is very low, the binary population would have few binary
products and be similar to a simple, single stellar population. The
investigation of the properties of binaries in 13 low-density GCs
revealed that the fraction of binaries ranges from 0.1 to 0.5 in the
core depending on the cluster \citep{Sollima2007}. Moreover, the
stellar density in GCs is sufficiently high to affect the binaries
that can be destroyed, created or modified by dynamical interaction
\citep[e.g.][]{Hut1992}. Higher stellar densities could result in
more efficient binary dissolution and thus a lower binary fraction
\citep{Marks2011}. Consequently, the binary population might have
different distributions and different fractions of binary products
in various clusters. A more detailed study of the binary population
with binary interactions might help to understand the dependence of
binary interactions on the cluster properties in the further
studies.
It is clear that our model is still quite simple and does not
investigate in much detail so far. In our investigation, we do not
consider the effect of the helium-rich core of the original
components on the subsequent evolution of binary products. Our
simulation shows that 36 per cent of all rapidly rotating stars are
formed by binary merger (including contact merger channel and
unstable RLOF merger channel). These rapidly rotating stars formed
by binary merger might contain the helium-rich core of the original
primary, and their remaining relative lifetimes are smaller than
those of single stars with the same masses
\citep[e.g.][]{Sills1997,Glebbeek2008,deMink2013}. For rapidly
rotating stars produced through stable RLOF, this effect need not to
be considered because they are the less-massive components of
binaries and their cores are still hydrogen rich at the onset of
mass transfer \citep[e.g.][]{Sills1997,Glebbeek2008,deMink2013}. For
rotating velocity, we simply assume that all binary products rotate
initially at 80\% of break-up velocity. A lower initial rotating
velocity or a faster spin-down will lead to a weaker variation of
surface abundances. These effects need more detailed investigation
in further study. In addition, the observations suggested that the
fraction of binaries in GCs may be lower than that in the field and
open clusters \citep{Rubenstein1997}. This might be because GCs are
much older than the field and open clusters, and consequently more
binaries in GCs are not expected to be recognized as binaries due to
their large luminosity ratio and their large orbital periods, or
they may not even be binaries due to merger \citep{deMink2013}.
\acknowledgments
We thank an anonymous referee for his/her valuable comments. This
work was supported by the Natural Science Foundation of China (Nos
11073049, 11033008, 11103073, 11390374 and 11373063), Science and
Technology Innovation Talent Programme of the Yunnnan Province (No.
2013HA005) and the Western Light Youth Project.
|
\section{Conclusion}
\label{sec:conclusion}
We showed that if we want to solve $\tgp(G,T)$ with a finite guard candidate set $G$, we can find a finite witness set $W(G)$, such that a solution of $\tgp(G,W(G))$ is feasible for $\tgp(G,T)$.
Our main result is the construction of a finite set of guard candidates $U$ of size $\bigO(n^2)$ such that $U$ admits an optimal cover of $T$.
Combining the two discretizations we concluded that an optimal solution of $\tgp(U, W(U))$ is also an optimal solution of $\tgp(T,T)$.
A polynomial time approximation scheme (PTAS) for $\tgp(T,T)$ using a former PTAS~\cite{gkkv-aasftg-09} for the discrete terrain guarding problem $\tgp(G,W)$ immediately follows.
Moreover, we formulate $\tgp(T,T)$ as an integer program (IP), yielding exact solutions.
\section{Discretization}
\label{sec:discretization}
This section is our core contribution.
We consider the following problem:
Given a terrain $T$, construct finite sets $G, W \subset T$ (guard candidate and witness points), such that any optimal (minimum-cardinality) solution for $\tgp(G,W)$ is optimal for $\tgp(T,T)$ as well.
We achieve this in three steps.
(1)~Provided with some finite guard candidate set $G \subset T$, Section~\ref{sec:discrete-witnesses} shows how to construct a finite witness set $W(G)$ from it, such that any feasible solution of $\tgp(G, W(G))$ is feasible for $\tgp(G,T)$ as well.
(2)~Section~\ref{sec:discrete-guards} discusses a finite set of guards $U$ that allows minimum-cardinality coverage of $T$.
(3)~In Section~\ref{sec:compl}, we argue that $\tgp(T,T)$ can be optimally solved using $\tgp(U,W(U))$, which is useful for a PTAS (Section~\ref{sec:ptas}) as well as for exact solutions (Section~\ref{sec:exact}).
Note that a discretization similar to that of Chwa et~al.\ for polygons~\cite{cjkmos-gagbgw-05} does not work.
Chwa et~al.\ pursue the idea of \emph{witnessable} polygons, which allow placing a finite set of witnesses, such that covering the witnesses implies full coverage of the polygon.
They show that this is possible if the polygon can be covered by a finite set of visibility kernels, \ie, if no point has a point-shaped visibility kernel.
Unfortunately, we can easily construct a terrain $T$ not coverable by a \emph{finite} union of \emph{visibility kernels}
$\VK(w) = \{ w' \in T \mid \V(w) \subseteq \V(w') \}$,
\ie, $T$ with finite $\left| \VK(w) \right|$ for some $w \in T$, compare Figure~\ref{fig:vis-kernel}.
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{ipe/vis_kernel.pdf}
\end{center}
\caption{Witness $w$, $\V(w)$ highlighted in red, and finite visibility kernel $\VK(w) = \{w, w', w''\}$ marked in blue.}
\label{fig:vis-kernel}
\end{figure}
\subsection{Witnesses}
\label{sec:discrete-witnesses}
Suppose we are given a terrain $T$ and a finite set $G \subset T$ of guard candidates
with $\V(G) = T$ and we want to cover $T$ using only guards $C \subseteq G$, \ie, we want to solve $\tgp(G,T)$.
$G$ could be the set $V$ of vertices to solve the vertex guard variant of the TGP or any other finite set,
especially the one in Equation~\eqref{eq:u} of Section~\ref{sec:discrete-guards},
which contains all guard candidates necessary to find an optimal solution of the
continuous version of the problem, $\tgp(T,T)$.
$G$ is finite by assumption, but $T$ isn't, so we generate a finite set $W(G) \subset T$ of witness points, such that any feasible solution for $\tgp(G,W(G))$ also is feasible for $\tgp(G,T)$.
Let $g \in G$ be one of the guard candidates.
$\V(g)$ subdivides $T$ into $\bigO(n)$ closed subterrains.
For the sake of simplicity, we project those subterrains onto
the $x$-axis, allowing us to represent $\V(g)$ as a set of closed intervals.
We consider the overlay of all visibility intervals of all guard candidates in $G$, see Figure~\ref{fig:overlay} for an overlay of two guards.
It forms a subdivision consisting of \emph{maximal intervals} and \emph{end points}.
Every point in a \emph{feature} $f$ (either end point or maximal interval) of the subdivision is seen by the same set of guards
\begin{equation}
G(f) = \left\{ g \in G \mid f \subseteq \V(g) \right\}.
\end{equation}
It is possible to simply place one witness in every feature of the subdivision.
Covering all $\bigO(n \cdot |G|)$ witnesses implies full coverage of $T$.
Similar to the shadow atomic visibility polygons in~\cite{crs-aaafmvgoag-11},
we can further reduce the number of witnesses by only using those
features $f$ with inclusion-minimal $G(f)$:
\todo{Maybe replace by survey cite later}
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{ipe/overlay.pdf}
\end{center}
\caption{Visibility overlay of two guards $g_1$ and $g_2$.
$\V(g_1)$ is indicated in blue; $\V(g_2)$ in
red; $\V(g_1) \cap $$\V(g_2)$ in orange. }
\label{fig:overlay}
\end{figure}
\begin{theorem}\label{thm:w}
Consider a terrain $T$ and a finite set of guard candidates $G$ with $\V(G) = T$.
Let $F$ denote the features of the visibility overlay of $G$ and $w_f \in f$ an arbitrary point in the feature $f \in F$.
Then for
\begin{equation}\label{eq:w}
W(G) = \left\{ w_f \mid f \in F, \, \textnormal{$G(f)$ is inclusion-minimal} \right\}
\end{equation}
any feasible solution of $\tgp(G, W(G))$ is feasible for $\tgp(G,T)$.
\end{theorem}
\begin{proof}
Let $C \subseteq G$ be a feasible cover of $W(G)$ and suppose some
point $w \in T$ is not covered by $C$.
By assumption, some point in $G$ can see $w$, so $w$ must be part
of some feature $f$ of the visibility overlay of $G$.
$W(G)$ either contains some witness in $w_f \in f$ or a witness
$w_{f'}$ with $G(f') \subseteq G(f)$ by construction.
In the first case, $w$ must be covered, otherwise $w_f$ would
not be covered and $C$ would be infeasible for $\tgp(G,W(G))$.
In the second case $w_{f'}$ is covered, so some guard in $G(f')$
is part of $C$, but that guard also covers $f$ and therefore $w$, a contradiction.
\end{proof}
\begin{figure*}
\begin{center}
\includegraphics[width=0.7\textwidth]{ipe/inclusion_minimal}
\end{center}
\caption{
The set of inclusion-minimal features may still be $\bigO(n \cdot |G|)$.
For $n\in \bigO(|G|)$ consider the terrain with $\bigO(n)$ valleys, with
$|G|/3$ guards placed on both slopes (blue on the left, red on the right).
In addition there is one black guard placed in each valley.
Thus, each of the $\bigO(n)$ valleys contains $\bigO(G)$ inclusion-minimal
intervals depicted in violet.
}
\label{fig:inclusion_minimal}
\end{figure*}
\begin{obs}\label{obs:witcard}
Using only the set of witnesses on inclusion-minimal features may not reduce the worst case complexity of $\bigO(n \cdot |G|)$ witnesses, see Figure~\ref{fig:inclusion_minimal}.
\end{obs}
Nevertheless, we expect inclusion-minimal witnesses to speed up an implementation.
\begin{obs}
$W(G)$ does not need to contain any interval end point.
An end point $p$ with adjacent maximal intervals
$I_1, I_2$ can always be left out, because $G(p) = G(I_1) \cup G(I_2)$.
\end{obs}
\subsection{Guard Positions}
\label{sec:discrete-guards}
Throughout this section, let $T$ be a terrain, $V$ its vertices and $E$ its edges;
let $C \subset T$ be some finite, not necessarily optimal, guard cover of $T$.
Moreover, let $U$ be the union of $V$ with all $x$-extremal points of all visibility
regions of all vertices:
\begin{equation}\label{eq:u}
U := V \cup \bigcup_{v \in V} \left\{ p \mid \textnormal{$p$ is extremal in $\V(v)$} \right\}.
\end{equation}
\begin{obs}\label{obs:guardcard}
The set $U$ has cardinality $\bigO(n^2)$, as noted by Ben-Moshe et~al.~\cite{bkm-acfaafotg-07}.
\end{obs}
Ben-Moshe et~al. also add an arbitrary point of $T$ between each pair of consecutive points in $U$.
They use this extended set as their witness set.
We, however, show in this section that $U$ is sufficient to admit an optimal guard cover for $T$.
Our basic idea is that for any cover $C$ it is always possible to move guards in $C\setminus U$ to a neighboring point in $U$ without losing coverage.
In particular, this is possible for an optimal guard cover.
First observe that we can not lose coverage for an edge $e$ that is entirely covered by a guard $g \in C \setminus U$ if we move $g$ to one of its neighbors in $U$.
\begin{lemma}\label{lem:guard-edge-u}
Let $g \in C \setminus U$ be a guard that covers an entire edge $e_i \in E$.
Then $u_\ell, u_r$, the ``$U$-neighbors'' of $g$ with
\begin{equation}\label{eq:ul-ur}
\begin{aligned}
u_\ell & = \max\{ u \in U \mid u < g \} \\
u_r & = \min\{ u \in U \mid g < u \}
\end{aligned}
\end{equation}
each entirely cover $e_i$, too.
\end{lemma}
\begin{proof}
$g$ covers $e_i$, so $v_i, v_{i+1} \in \V(g)$, implying $g \in \V(v_i) \cap \V(v_{i+1})$.
Moving $g$ towards $u_\ell$ does not move $g$ out of $\V(v_i)$ or $\V(v_{i+1})$, as the boundaries of those regions are contained in $U$ by construction.
So $v_i, v_{i+1} \in \V(u_\ell)$ and $e_i \subseteq \V(u_\ell)$.
Analogously: $e_i \subseteq \V(u_r)$
\end{proof}
It remains to consider the \emph{critical edges}, \ie, those that are not entirely covered by a single guard, compare Figure~\ref{fig:critical_edge}.
\begin{figure}[b]
\begin{center}
\includegraphics[width=0.45\textwidth]{ipe/critical_edge.pdf}
\end{center}
\caption{The edge $e$ is critical w.r.t. the given guard cover.
The right (left) part of $e_i$,
indicated in blue (red), is seen by $g_\ell$ ($g_r$) only.
}
\label{fig:critical_edge}
\end{figure}
\begin{definition}[Critical Edge]
$e \in E$ is a \emph{critical edge \wrt $g $} in the cover $C$ if
$C \setminus \{ g \}$ covers some part of, but not all of, $e$.
\end{definition}
That is, after removing $g$, $e$ is only partially covered.
\begin{definition}[Left-Guard/Right-Guard]
$g \in C$ is a \emph{left-guard (right-guard)} of
$e_i \in E$ if $g < v_i$ ($v_{i+1} < g$) and $e_i$ is critical \wrt $g$.
We call $g$ \emph{left-guard (right-guard)} if it is
a left-guard (right-guard) of some $e \in E$.
\end{definition}
For the sake of completeness, we state and prove the following lemma, which also follows from the well-established order claim~\cite{bkm-acfaafotg-07}:
\begin{lemma}\label{lem:single-interval-vis}
Let $g \in C$ be a left-guard (right-guard) of $e_i \in E$.
Then $g$ covers a single interval of $e_i$, which includes $v_{i+1}$ ($v_i$).
\end{lemma}
\begin{proof}
Refer to Figure~\ref{fig:critical_edge}.
Obviously, $g$ is nowhere below the line supporting $e_i$.
Let $p$ be a point on $e_i$ seen by $g$.
It follows that $\overline{g p}$ and $\overline{p v_{i+1}}$
form an $x$-monotone convex chain that is nowhere below $T$.
Thus, the secant $\overline{g v_{i+1}}$ is nowhere below $T$.
It follows that $g$ sees $v_{i+1}$ as well as any
point between $p$ and $v_{i+1}$ (same argument). The argument for the right-guard is analogous.
\end{proof}
\begin{cor}\label{cor:unique-left-guard}
For each critical edge $e$ there is exactly one
left-guard (right-guard) in $C$.
\end{cor}
\begin{cor}\label{cor:vis-intersect}
Let $e\in E$ be a critical edge and $g_\ell, g_r \in C$ be its
left- and right-guard. Then $\V(g_\ell)\cap e \cap \V(g_r) \not = \emptyset $.
\end{cor}
The following Lemma shows that we can move a guard $g \in C \setminus U$
to its left neighbor in $U$ without losing coverage of $T$ if
$g$ is not a right-guard.
\begin{lemma}\label{lem:left-guard-to-u}
Let $C$ be some finite cover of $T$, $g \in C \setminus U$
be a left- but no right-guard, and let $u_\ell$ be the
left $U$-neighbor of $g$ as in Equation~\eqref{eq:ul-ur}.
Then
\begin{equation}
C' = \left( C \setminus \{ g \} \right) \cup \{ u_\ell \}
\end{equation}
is a guard cover of $T$.
\end{lemma}
\begin{proof}
By Lemma~\ref{lem:guard-edge-u}, edges entirely covered by $g$ are also covered by $u_\ell$.
So consider $p \in \V(g)\cap e_r$ on a critical edge $e_r$ \wrt $g$ as
depicted in Figure~\ref{fig:left-guard-to-u}.
The guard $g$ is in the interior of an edge $e$ since $g \in C \setminus U$.
As $p$ is seen by $g$ it must be nowhere below the line supporting $e$.
It follows that segments $\overline{u_\ell g}$ and $\overline{gp}$ form an
$x$-monotone convex chain that is nowhere below $T$.
Hence, the secant $\overline{u_\ell p}$ is nowhere below $T$, so $u_\ell$ sees $p$.
In particular, it holds that
\begin{equation}
\V(g)\cap e_r \subseteq \V(u_\ell)\cap e_r,
\end{equation}
for every critical edge $e_r$ \wrt $g$.
\end{proof}
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{ipe/move-left.pdf}
\end{center}
\caption{Moving the left-guard $g$ further to the left.
Any point $p$ on critical edge $e_r$ that is seen by $g$
remains visible while moving $g$ to its left U-neighbor $u_\ell$.
Also, non-critical edge $e'$ remains entirely visible since $g$
does not cross $u_\ell$ which is induced by $v$ and $v'$.}
\label{fig:left-guard-to-u}
\end{figure}
\begin{cor}\label{cor:right-guard-to-u}
Let $C$ be some finite cover of $T$, $g \in C \setminus U$
be a right- but no left-guard, and let $u_r$ be the
right $U$-neighbor of $g$ as in Equation~\eqref{eq:ul-ur}.
Then
\begin{equation}
C' = \left( C \setminus \{ g \} \right) \cup \{ u_r \}
\end{equation}
is a guard cover of $T$.
\end{cor}
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{ipe/left-right.pdf}
\end{center}
\caption{No guard in $T \setminus U$ is left- and right-guard. Any point on the critical edge
$e_\ell$ that is seen by $g$ is also seen by the guard $g_r$.
Hence, $e_\ell$ can not be critical with respect to $g$, contradiction.
A symmetric argument applies for $e_r$.}
\label{fig:leftright-guard-to-u}
\end{figure}
\begin{lemma}\label{lem:leftright-guard-to-u}
Let $C$ be some finite cover of $T$.
No $g \in C \setminus U$ is both a left- and a right-guard.
\end{lemma}
\iffalse
\todo[inline]{
Alternative: Let $C$ be some finite cover of $T$.
A right-guard $g \in C \setminus U$ can not also be a left-guard. - bin fuer die, die drin ist, sonst wieder korollar
}
\fi
\begin{proof}
Refer to Figure~\ref{fig:leftright-guard-to-u}.
We prove the claim by contradiction.
Suppose that $g \in C \setminus U$ is
the left-guard for $e_r$ (to the right of $g$) and
the right-guard for $e_\ell$ (to the left of~$g$).
Since $g$ is the left-guard for critical edge $e_r$
there must also be the right-guard $g_r$ for $e_r$.
By Corollary~\ref{cor:vis-intersect} there is at least one point
$p_r\in e_r$ that is seen by $g$ and $g_r$.
Since $g \in C \setminus U$, it must be in the interior of some edge $e$.
Now consider $p_r$ and any point $p_\ell \in \V(g)$.
Both points are not below the line supported by $e$ and
the same holds for $g$ and $g_r$ with respect to $e_r$.
It follows that segments $\overline{p_\ell g}$, $\overline{gp_r}$,
and $\overline{p_r g_r}$ form an $x$-monotone convex chain that
is nowhere below $T$. Hence, $g_r$ sees $p_\ell$.
Thus, any point $p\in\V(g)$ to the left of $g$
is also seen by $g_r$, a contradiction to $g$ being a right-guard.
\end{proof}
The next theorem shows that the set $U$ as defined in Equation~\eqref{eq:u} contains all guard candidates necessary for a minimum-cardinality guard cover of $T$.
So even if we are allowed to place guards anywhere on $T$, we only need those in $U$ and thus have discretized the problem.
\begin{theorem}\label{thm:g}
Let $T$ be a terrain, $C \subset T$ a finite guard cover of $T$, possibly of minimum cardinality, and consider $U$ as defined in Equation~\eqref{eq:u}.
Then there exists a guard cover $C' \subseteq U$ of $T$ with $|C'| = |C|$.
\end{theorem}
\begin{proof}
We iteratively replace a guard $g \in C \setminus U$ by one in $U$ until $C \subseteq U$.
This maintains the cardinality of $C$, thus constructing the set $C'$ as claimed.
Should $g$ be neither left- nor right-guard, it can be moved to a neighboring point in $U$ by Lemma~\ref{lem:single-interval-vis}.
If, on the other hand, $g$ is only a left-, but not a right-guard (or vice versa), it can be moved to its left (right) neighbor in $U$ as shown in Lemma~\ref{lem:left-guard-to-u} and Corollary~\ref{cor:right-guard-to-u}.
Lemma~\ref{lem:leftright-guard-to-u} states that $g$ cannot be a left- and a right-guard at the same time.
\end{proof}
\subsection{Complete Discretization}
\label{sec:compl}
In this section, we formulate our key result.
Let $\opt(G,W)$ denote the cardinality of an optimal, \ie, minimum-cardinality, solution for $\tgp(G,W)$.
\begin{theorem}\label{th:opt}
Let $T$ be a terrain, $U$ and $W(U)$ as defined in Equations~\eqref{eq:u} and~\eqref{eq:w}.
Then:
If $C$ is an optimal solution of $\tgp(U, W(U))$, \ie, $|C|=\opt(U,W(U))$, then $C$ is also an optimal solution of $\tgp(T,T)$, \ie, $\opt(T,T) = |C| = \opt(U,W(U))$.
\end{theorem}
\begin{proof}
We have $\opt(T,T) \leq \opt(U,T)$.
Theorem~\ref{thm:g} states that for an optimal guard cover $C''$ of $\tgp(T,T)$ there exists a guard cover $C'\subseteq U$ with $|C'| = |C''|$, \ie,
\begin{equation}
\opt(T,T) = |C''| = |C'| \geq \opt(U,T).
\end{equation}
This yields $\opt(T,T) = \opt(U,T)$.
If we are given an optimal solution $C$ to $\tgp(U,W(U))$, with $\opt(U, W(U))$ guards, $C$ is a feasible solution for $\tgp(U,T)$ according to Theorem~\ref{thm:w}. As $|C| = \opt(U,W(U)) \leq \opt(U,T)$ and $C$ is feasible for $\tgp(U,T)$, we have $\opt(U,W(U)) = \opt(U,T)$ which concludes the proof.
\end{proof}
\begin{obs}\label{obs:card}
Observations~\ref{obs:witcard} and~\ref{obs:guardcard} yield:
The set of guard candidates $U$ has cardinality $\bigO(n^2)$, the finite witness set $W(U)$ has cardinality $\bigO(n^3)$.
\end{obs}
\section{Exact Solutions}
\label{sec:exact}
Let $T$ be a terrain, and $U$ and $W(U)$ be defined as in Equations~\eqref{eq:u} and~\eqref{eq:w}.
Our discretization allows us to formulate the following integer program (IP)
to find an exact solution of $\tgp(T,T)$ by modeling guard candidates as binary variables and witnesses as constraints:
\begin{alignat}{3}
\textnormal{min} & \sum_{g \in U} x_g \label{eq:ip-begin} \\
\textnormal{s.\,t.} & \sum_{g \in \V(w) \cap U} x_g \geq 1 & \quad & \forall w \in W(U) \\
& x_g \in \{0, 1\} & \quad & \forall g \in U \label{eq:ip-end}
\end{alignat}
The IP in \eqref{eq:ip-begin}~--~\eqref{eq:ip-end} paves the way for an exact, while not polytime, algorithm for the continuous TGP.
\section{Introduction}
\label{sec:introduction}
Let a \emph{terrain} $T$ denote an $x$-monotone chain defined by its vertices $V = \{ v_1, \dots, v_n \}$.
It has edges $E = \{ e_1, \dots, e_{n-1} \}$ with $e_i = \overline{v_i v_{i+1}}$.
Due to its monotonicity the points on $T$ are totally ordered with regard to their $x$-coordinate.
For $p,q \in T$, we write $p < q$ if $p$ is \emph{left} of $q$, \ie, if $p$ has a smaller $x$-coordinate than $q$.
A point $p \in T$ \emph{sees} or \emph{covers} $q \in T$ iff $\overline{pq}$ is nowhere below $T$.
$\V(p)$ is the \emph{visibility region} of $p$ with $\V(p) = \{ q \in T \mid \textnormal{$p$ sees $q$} \}$.
$\V(p)$ is not necessarily connected, and can be considered as the union of $\bigO(n)$ maximal subterrains, compare Figure~\ref{fig:visibility_region}.
We say that $q \in \V(p)$ is \emph{extremal} in $\V(p)$, if $q$ has a maximal
or minimal $x$-coordinate within its connected component of $\V(p)$.
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{ipe/visibility_region.pdf}
\end{center}
\caption{The visibility region $\V(p)$ (blue) of point $p$.}
\label{fig:visibility_region}
\end{figure}
For $G \subseteq T$ we abbreviate $\V(G) := \bigcup_{g \in G} \V(g)$.
A set $G \subseteq T$ with $\V(G) = T$ is named a \emph{(guard) cover} of $T$.
In this context, $g \in G$ is referred to as \emph{guard}.
\begin{definition}[Terrain Guarding Problem]\label{def:tgp}
In the \emph{Terrain Guarding Problem (TGP)}, we are given a terrain $T$ and sets of guard candidates and witnesses $G, W \subseteq T$.
We seek a minimum-cardinality guard cover $G^* \subseteq G$ such that $W \subseteq \V(G^*)$, \ie, $G^*$ covers $W$, and abbreviate this problem by $\tgp(G,W)$.
We assume $W \subseteq \V(G)$, \ie, that $\tgp(G,W)$ always has a feasible solution.
\end{definition}
Note that $\tgp(T,T)$ is the continuous TGP and $\tgp(V,T)$ is the special case with vertex guards.
\input{rw}
\subsection{Our Contribution}
\label{sec:contribution}
We present a discretization, \ie, finite sets $G, W \subset T$ such that any optimal solution for $\tgp(G,W)$
is optimal and feasible for $\tgp(T,T)$ as well:
\begin{enumerate}
\item
For the sake of completeness we argue that for each finite guard candidate set $G$ there exists a finite witness set $W(G)$, such that a solution for $\tgp(G,W(G))$ is feasible for $\tgp(G,T)$ (Section~\ref{sec:discrete-witnesses}).
\item
For each terrain $T$ there is a finite guard candidate set $U$, such that for each (possibly optimal) guard cover $C \subset T$ for $\tgp(T,T)$ there exists a guard cover $C' \subseteq U$ with $|C'| = |C|$ (Section~\ref{sec:discrete-guards}).
\item
It then follows that any feasible optimal solution of $\tgp(U,W(U))$ is also optimal and feasible for $\tgp(T,T)$ (Section~\ref{sec:compl}).
\item
Combining this discretization with the PTAS of Gibson et al.~\cite{gkkv-aasftg-09} for the discrete $\tgp(G,W)$ case, we obtain a PTAS for $\tgp(T,T)$ (Section~\ref{sec:ptas}).
\item
The discretization also yields an IP formulation for exact solutions (Section~\ref{sec:exact}).
\end{enumerate}
\section{The PTAS}
\label{sec:ptas}
In this section we combine the PTAS of Gibson et al.~\cite{gkkv-aasftg-09}
with the results from Section~\ref{sec:discretization} to obtain a PTAS for $\tgp(T,T)$.
Let us first formulate the result of Gibson et al., who presented a PTAS for the discrete terrain guarding problem, in our notation:
\begin{lemma}[PTAS by Gibson et al.~\cite{gkkv-aasftg-09}]\label{le:gibson}
Let $T$ be a terrain, and $G,W \subset T$ finite sets of guard candidates and points to be guarded. Then there exists a polynomial time approximation scheme for $\tgp(G,W)$. That is, there exists a polynomial time algorithm that returns a subset $C\subseteq G$ with $W\subseteq \V(C)$, such that $|C| \leq (1+\epsilon)\cdot \opt(G,W) \; \forall \epsilon > 0$, where $\opt(G,W)$ denotes the optimal solution for $\tgp(G,W)$.
\end{lemma}
We can now easily combine Theorem~\ref{th:opt} and Lemma~\ref{le:gibson} for a PTAS for the continuous TGP:
\begin{theorem}\label{th:ptas}
Let $T$ be a terrain. Then there exists a polynomial time approximation scheme for $\tgp(T,T)$, the continuous terrain guarding problem. That is, there exists a polynomial time algorithm that returns a subset $C\subset T$ with $\V(C)=T$, such that $|C| \leq (1+\epsilon)\cdot \opt(T,T) \; \forall \epsilon > 0$, where $\opt(T,T)$ denotes the optimal solution for $\tgp(T,T)$.
\end{theorem}
\begin{proof}
Using Equations~\eqref{eq:u} and~\eqref{eq:w} we determine the sets $U$ and $W(U)$ for the terrain $T$, two finite subsets of $T$. Given an arbitrary $\epsilon > 0$ we can compute a set $C\subseteq U \subset T$ with $\V(C)=T$ such that:
\begin{equation}\label{eq:eps}
|C| \leq (1+\epsilon)\cdot \opt(U,W(U))
\end{equation}
using the PTAS of Gibson et al., Lemma~\ref{le:gibson}.
Moreover, Theorem~\ref{th:opt} yields:
\begin{equation}\label{eq:id}
\opt(U,W(U)) = \opt(T,T).
\end{equation}
Combining Equations~\eqref{eq:eps} and~\eqref{eq:id} we obtain
\begin{equation}
\begin{aligned}
|C| & \leq (1 + \epsilon) \cdot \opt(U, W(U)) \\
& = (1 + \epsilon) \cdot \opt(T, T)
\end{aligned}
\end{equation}
as claimed.
\end{proof}
\subsection{Related Work}
\label{sec:rw}
The terrain guarding problem is closely related to the well known \emph{Art Gallery Problem} (AGP) where the objective is to find a minimum cardinality guard set that allows complete coverage of a polygon $P$.
Many different versions of this problem have been considered, including variants with guards restricted to be located on vertices (\emph{vertex guards}), patrolling along edges (\emph{edge guards}), or located on arbitrary positions in $P$ (\emph{point guards}); each in simple polygons and in polygons with holes.
The first result was obtained by Chv\'{a}tal~\cite{c-actpg-75}, who proved the ``Art Gallery Theorem'', answering a question posed by Victor Klee in 1973: $\lfloor \frac{n}{3} \rfloor$ many guards are always sufficient and sometimes necessary to guard a polygon of $n$ vertices.
A simpler and elegant proof of the sufficiency was later given by Fisk~\cite{f-spcwt-78}.
Related results were obtained for different polygon classes.
The optimization problem was shown to be NP-hard for various problem versions~\cite{os-snpdp-83,sh-tnhag-95}, even the allegedly easier problem of finding a minimum cardinality vertex guard set in simple polygons is NP-hard~\cite{ll-ccagp-86}.
Eidenbenz et al.~\cite{esw-irgpt-01} gave bounds on the approximation ratio: For polygons with holes a lower bound of $\Omega(\log n)$ holds, for vertex, edge and point guards in simple polygons they showed the problem to be APX-hard.
For detailed surveys on the AGP see O'Rourke~\cite{r-agta-87} or Shermer~\cite{s-rrag-92} for classical results.
\todo[inline]{or~\cite{} for more recent computational developments.}
Motivation for terrain guarding is the placement of street lights or security cameras along roads~\cite{gkkv-aasftg-09}, or the optimal placement of antennas for line-of-sight communication networks~\cite{bkm-acfaafotg-07}.
For the terrain guarding problem the focus was on approximation algorithms, because NP-hardness was generally assumed, but had not been established by then.
The first who were able to establish a constant-factor approximation algorithm were Ben-Moshe et al.~\cite{bkm-acfaafotg-07}.
They presented a combinatorial constant-factor approximation for the discrete vertex guard problem version $\tgp(V,V)$, where only vertex guards are used to cover only the vertices, and were able to use it as a building block for an $\bigO(1)$-approximation of the continuous terrain guarding variant $\tgp(T,T)$.
The approximation factor of this algorithm was never stated by the authors, but was claimed to be 6 in~\cite{k-a4aafgdt-06} (with minor modifications).
Another constant-factor approximation based on $\epsilon$-nets and Set Cover was given by Clarkson and Varadarajan~\cite{cv-iaags-07}.
King~\cite{k-a4aafgdt-06} presented a 4-approximation (which was later shown to actually be a 5-approximation~\cite{k-eaag}), both for the discrete $\tgp(V,V)$ and the continuous $\tgp(T,T)$ problem.
The most recent $\bigO(1)$-approximation was presented by Elbassioni et al.~\cite{emms-iafgdt-08,ekmms-iag15-11}:
For non-overlapping discrete sets $G,W \subset T$ LP-rounding techniques lead to a 4-approximation
(5-approximation if $G \cap W \neq \emptyset$) for $\tgp(G,W)$ as well as for
the continuous case $\tgp(T,T)$.
This approximation is also applicable for the more general \emph{weighted} terrain guarding problem:
Weights are assigned to the guards and a minimum weight guard set is to be identified.
Finally, in 2009, Gibson et al.~\cite{gkkv-aasftg-09} showed that the discrete terrain guarding problem allows a \emph{polynomial time approximation scheme} (PTAS) based on local search.
They present PTASs for two problem variants: for $\tgp(G,W)$ where $G$ and $W$ are (not necessarily disjoint) finite subsets of the terrain $T$ and for $\tgp(G,T)$, \ie, the variant with a finite guard candidate set $G$.
For the continuous case, \ie, $\tgp(T,T)$, they claim that the local search works as well, but that they were not yet able to limit the number of bits needed to represent the guards maintained by the local search.
Thus, to the best of our knowledge, no PTAS for $\tgp(T,T)$ has been established until now.
The NP-hardness of the TGP was settled after all these approximation results: King and Krohn~\cite{kk-tginph-10} proved both the discrete and the continuous case to be NP-hard by a reduction from PLANAR 3SAT in 2006.
Other problems considered in the context of terrains include, for example, guards that are allowed to ``hover'' over the terrain~\cite{e-aatg-02}, or the computation of visibility polygons, \ie, the set of points on the terrain visible to a point $p$ on the terrain~\cite{lss-facvm-14}.
|
\section{Introduction}
\label{paperE:sec:introduction}
Linear support vector machine (SVM) has become the classifier of choice for many large scale classification problems.
The main reasons for the success of linear SVM are its max margin property achieved through a convex optimization, a training time linear in the size of the training data, and a testing time independent of it.
Although the linear classifier operating on the input space is usually not very flexible, a linear classifier operating on a mapping of the data to a higher dimensional feature space can become arbitrarily complex.
Mixtures of linear classifiers has been proposed to increase the non-linearity of linear classifiers \cite{mixing_svm,AghazadehASC12}; which can be seen as feature mappings augmented with non-linear gating functions.
The training of these mixture models usually scales bilinearly with respect to the data and the number of mixtures.
The drawback is the non-convexity of the optimization procedures, and the need to know the (maximum) number of components beforehand.
Kernelized SVM maps the data to a possibly higher dimensional feature space, maintains the convexity, and can become arbitrarily flexible depending on the choice of the kernel function.
The use of kernels, however, is limiting.
Firstly, kernelized SVM has significantly higher training and test time complexities when compared to linear SVM.
As the number of support vectors grows approximately linearly with the training data \cite{Steinwart04sparsenessof},
the training complexity becomes approximately somwehere between $\mathcal{O}(n^2)$ and $\mathcal{O}(n^3)$.
Testing time complexity scales linearly with the number of support vectors,
bounded by $\mathcal{O}(n)$.
Secondly, the positive (semi) definite (PSD) kernels are sometimes not expressive enough to model various sources of variation in the data.
A recent study \cite{good_rec_nonmetric} argues that metric constraints are not necessarily optimal for recognition.
For example, in image classification problems, considering kernels as similarity measures, they cannot align exemplars, or model deformations when measuring similarities.
As a response to this, invariant kernels were introduced \cite{invariant_kernels} which are generally indefinite.
Indefinite similarity measures plugged in SVM solvers result in non-convex optimizations, unless explicitly made PSD, mainly using eigen decomposition methods \cite{sim_class}.
Alternatively, latent variable models have been proposed to address the alignment problem \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot \cite{FelzenszwalbGMR10,YangWVM12}.
In these cases, the dependency of the latent variables on the parameters of the model being learnt mainly has two drawbacks:
1) the optimization problem in such cases becomes
non-convex,
and 2) the cost of training becomes much higher than the case without the latent variables.
This paper aims to address these problems using explicit basis expansion.
We show that the resulting model:
1) has better training and test time complexities than kernelized SVM models,
2) can make use of indefinite similarity measures without any need for removal of the negative eigenvalues, which requires the expensive eigen decomposition,
3) can make use of multiple similarity measures without losing convexity, and with a cost linear in the number of similarity measures.
Our contributions are: 1) proposing and analyzing Basis Expanding SVM (BE-SVM) regarding the aforementioned three properties, and 2) investigating the suitability of particular forms of invariant similarity measures for large scale visual recognition problems.
\section{Basis Expanding Support Vector Machine}
\label{paperE:sec:model}
\subsection{Background: SVM}
\label{paperE:sec:model:background}
Given a dataset $\mathcal{D}=\{(x_1,y_1), \ldots, (x_n,y_n)| x_i \in \mathcal{X},\, y_i \in \{-1,1\}\}$ the SVM based classifiers learn max margin binary classifiers.
The SVM classifier is $ f(x)=\langle w, x\rangle\geq 0$ \footnote{We omit the bias term for the sake of clarity.}.
The $w$ is learnt via minimizing $\frac{1}{2}\langle w,w \rangle + C\sum_i \ell_H(y_i, f(x))$, where $\ell_H(y,x) = \max(0,1-xy)$ is the Hinge loss.
Any positive semi definite (PSD) kernel $k:\mathcal{X}\times\mathcal{X}\rightarrow\mathbb{R}$ can be associated with a reproducing kernel hilbert space (RKHS) $\mathcal{H}$, and vice versa, that is
$\langle \psi_\mathcal{H}(x),\psi_\mathcal{H}(y) \rangle = k(x,y)$,
where $\psi_{\mathcal{H}}:\mathcal{X}\rightarrow \mathcal{H}$ is the implicitly defined feature mapping associated to $\mathcal{H}$ and consequently to $k(.,.)$.
Representer theorem states that in such a case,
$\psi_\mathcal{H}(w)=\sum_i \gamma_i k(.,x_i)$ where $\gamma_i \in \mathbb{R}\,\, \forall i$.
For a particular case of $k(.,.)$, namely the linear kernel $k(\mathbf{x},\mathbf{y})=\mathbf{x}\cdot\mathbf{y}$ associated with an Euclidean space, linear SVM classifier is
~$f_l(\mathbf{x})=\mathbf{w}^\text{T}\mathbf{x}\geq 0$
where $\mathbf{w}$ is given by minimizing the primal SVM objective:
~$\frac{1}{2}\|\mathbf{w}\|^2 + C\sum_i \ell_H(y_i,f_l(\mathbf{x}_i))$.
More generally, given an arbitrary PSD kernel $k(.,.)$, the kernelized SVM classifier is
~$ f_k(x) = \sum_i \alpha_i k(x,x_i) \geq 0 $
where $\alpha_i$s are learnt by minimizing the dual SVM objective:
~$\frac{1}{2}\mathbf{\alpha}^\text{T}\mathbf{Y}\mathbf{K}\mathbf{Y}\mathbf{\alpha}- \|\mathbf{\alpha}\|_1,\, 0\leq \alpha_i \leq C, \, \mathbf{\alpha}^\text{T}\mathbf{y} = 0$
where $\mathbf{Y}=\diag{\mathbf{y}}$.
The need for positiveness of $k(.,.)$ is evident
in the dual SVM objective
where the quadratic regularizing term depends on the eigenvalues of $\mathbf{K}_{ij} = k(x_i,x_j)$.
In case of indefinite $k(.,.)$s, the problem becomes non-convex and the inner products need to be re-defined, as there will be no associating RKHS to indefinite similarity measures.
Various workarounds for indefinite similarity measures exist, most of which involve expensive eigen decomposition of the gram matrix \cite{sim_class}.
A PSD kernel can be learnt from the similarity matrix, with some constraints \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot being close to the similarity matrix where closeness is usually measured by the Frobenius norm.
In case of Frobenius norm, the closed form solution is spectrum clipping, namely setting the negative eigenvalues of the gram matrix to 0 \cite{sim_class}.
As pointed out in \cite{learning_kernels_from_indef}, there is no guarantee that the resulting PSD kernels are optimal for classification.
Nevertheless, jointly optimizing for a PSD kernel and the classifier \cite{learning_kernels_from_indef} is impractical for large scale scenarios.
We do not go into the details of possible re-formulations regarding indefinite similarity measures, but refer the reader to \cite{ong04leanpk,Indefinite_Kernels,sim_class} for more information.
Linear and Kernelized SVM have very different properties.
Linear SVM has a training cost of $\mathcal{O}(d_\mathbf{x}n)$ and a testing cost of $\mathcal{O}(d_\mathbf{x})$ where $d_\mathbf{x}$ is the dimensionality of $\mathbf{x}$ .
Kernelized SVM has a training complexity which is $\mathcal{O}(d_k (n n_{sv}) + n_{sv}^3)$ \cite{Keerthi06buildingsupport} where $d_k$ is the cost of evaluating the kernel for one pair of data, and $n_{sv}$ is the number of resulting support vectors.
The testing cost of kernelized SVM is $\mathcal{O}(d_kn_{sv})$.
Therefore, a significant body of research has been dedicated to reducing the training and test costs of kernelized SVMs by approximating the original problem.
\subsection{Speeding up Kernelized SVM}
\label{paperE:sec:model:speedup}
A common approach for approximating the kernelized SVM problem is to restrict the feature mapping of $w$:
$\psi_\mathcal{H}(w) \approx \psi_R(w) = \sum_{j=1}^J \beta_j \psi_\mathcal{H} (z_j)$ where $J < n$.
Methods in this direction either learn synthetic samples $z_j$ \cite{direct_sparse} or restrict $z_j$ to be on the training data \cite{Keerthi06buildingsupport}.
These methods essentially exploit low rank approximations of the gram matrix $\mathbf{K}$.
Low rank approximations of PD $\mathbf{K} \succ 0$, result in speedups in training and testing complexities of kernelized SVM.
Methods that learn basis coordinates outside the training data \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot \cite{direct_sparse}, usually involve intermediate optimization overheads, and thus are prohibitive in large scale scenarios.
On the contrary, the Nystr\"{o}m method gives a low rank PSD approximation to $\mathbf{K}$ with a very low cost.
The Nystr\"{o}m method \cite{Williams00theeffect} approximates $\mathbf{K}$ using a randomly selected subset of the data:
\begin{equation}
\mathbf{K} \approx \mathbf{K}_{\mathbf{{n}}\mathbf{{m}}} \mathbf{K}_{\mathbf{{m}}\mathbf{{m}}} ^{-1} \mathbf{K}_{\mathbf{{m}}\mathbf{{n}}}
\label{paperE:eqn:nystrom_low_rank}
\end{equation}
where
$\mathbf{K}_{\mathbf{{a}}\mathbf{{b}}}$ refers to a sub matrix of $\mathbf{K} = \mathbf{K}_{\mathbf{{n}}\mathbf{{n}}}$ indexed by $\mathbf{{a}}=(a_1,\ldots,a_n)^\text{T}, \, a_i \in \{0,1\}$, and similarly by $\mathbf{{b}}$.
The approximation (\ref{paperE:eqn:nystrom_low_rank}) is derived by defining eigenfunctions of $k(.,.)$ as expansions of numerical eigenvectors of $\mathbf{K}$.
A consequence is that the data can be embedded in an Euclidean space:
$\mathbf{K} \approx \mathbf{\Psi}_{\mathbf{{m}}\mathbf{{n}}}^\text{T}\mathbf{\Psi}_{\mathbf{{m}}\mathbf{{n}}}$,
where $\mathbf{\Psi}_{\mathbf{{m}}\mathbf{{n}}}$, the Nystr\"{o}m feature space, is
\begin{equation}
\mathbf{\Psi}_{\mathbf{{m}}\mathbf{{n}}}=\mathbf{K}_{\mathbf{{m}}\mathbf{{m}}}^{-\frac{1}{2}}\mathbf{K}_{\mathbf{{m}}\mathbf{{n}}}
\label{paperE:eqn:nystrom_feature_map}
\end{equation}
Methods exist which either explicitly or implicitly exploit this \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot \cite{Lee01rsvm:reduced} to reduce both the training and test costs, by restricting the support vectors to be a subset of
the bases defined by $\mathbf{m}$.
In case of indefinite similarity measures, $\mathbf{K}_{\mathbf{{m}}\mathbf{{m}}}^{-\frac{1}{2}}$ in (\ref{paperE:eqn:nystrom_feature_map}) will not be real.
In the rest of the paper, we refer to an indefinite version of a similarity matrix $\mathbf{K}$ with $\tilde{\mathbf{K}}$, and refer to the normalization by $\mathbf{K}_{\mathbf{{m}}\mathbf{{m}}}^{-\frac{1}{2}}$ with Nystr\"{o}m normalization.
In order to get a PSD approximation of an indefinite $\tilde{\mathbf{K}}$, the indefinite $\tilde{\mathbf{K}}_{\mathbf{{m}}\mathbf{{m}}}$ (\ref{paperE:eqn:nystrom_low_rank}) needs to be made PSD.
Spectrum clipping, spectrum flip, spectrum shift, and spectrum square are possible solutions based on eigen decomposition of $\tilde{\mathbf{K}}_{\mathbf{{m}}\mathbf{{m}}}$.
The latter can be achieved without the eigen decomposition step: $\tilde{\mathbf{K}}_{\mathbf{{m}}\mathbf{{m}}}^\text{T}\tilde{\mathbf{K}}_{\mathbf{{m}}\mathbf{{m}}} \succeq 0$.
If the goal is to find the PSD matrix closest to the original indefinite $\tilde{\mathbf{K}}$ with respect to the reduced basis set $\mathbf{m}$, spectrum clip gives the closed form solution.
Therefore, when there are a few negative eigenvalues, the spectrum clip technique gives good low rank approximations to $\tilde{\mathbf{K}}_{\mathbf{{m}}\mathbf{{m}}}$ which can be used by (\ref{paperE:eqn:nystrom_low_rank}) to get a low rank PSD approximation to $\tilde{\mathbf{K}}$.
However, when there are a considerable number of negative eigenvalues, as it is the case with most of the similarity measures we consider later on in section \ref{paperE:sec:model:similarity_measures},
there is no guarantee for the resulting PSD matrix to be optimal for classification.
This is true specially when eigenvectors associated with negative eigenvalues contain discriminative information.
We experimentally verify in section \ref{paperE:sec:experiments:invariant} that the negative eigenvalues do contain discriminative information.
We seek normalizations that do not assume a PSD $\mathbf{K}_{\mathbf{m}\mathbf{m}}$, and do not require eigen-decompositions.
For example, one can replace $\mathbf{K}_{\mathbf{{m}}\mathbf{{m}}}$ in (\ref{paperE:eqn:nystrom_feature_map}) with the covariance of columns of $\mathbf{K}_{\mathbf{{m}}\mathbf{{n}}}$.
We experimentally found out that a simple embedding, presented in the next section in (\ref{paperE:eqn:besvm_feature_map_norm}),
is competitive with the Nystr\"{o}m embedding (\ref{paperE:eqn:nystrom_feature_map}) for PSD similarity measures,
while outperforming it in case of indefinite ones that we studied.
\subsection{Basis Expanding SVM}
\label{paperE:sec:model:be_svm}
Basis Expanding SVM (BE-SVM) is a linear SVM classifier equipped with a normalization of the following explicit feature map
\begin{equation}
\tilde\varphi(\mathbf{x}) = \left[ s(\mathbf{b}_1,\mathbf{x}), \ldots, s(\mathbf{b}_B,\mathbf{x})\right]^\text{T}
\label{paperE:eqn:besvm_feature_map}
\end{equation}
where $\mathcal{B} = \{\mathbf{b}_1,\ldots, \mathbf{b}_B\}$ is an ordered basis set\footnote{For the moment assume $\mathcal{B}$ is given. We experiment with different basis selection strategies in section \ref{paperE:sec:experiments:basis_sel}).} which is a subset of the training data, and $s(.,.)$ is a pairwise similarity measure.
The BE-SVM feature space defined by
\begin{equation}
\varphi(\mathbf{x}) = \frac{1}{\mathbb{E}_\mathcal{X}[\|\tilde\varphi - \mathbb{E}_\mathcal{X}[\tilde\varphi] \|]} \left(\tilde\varphi(\mathbf{x}) - \mathbb{E}_\mathcal{X}[\tilde\varphi] \right)
\label{paperE:eqn:besvm_feature_map_norm}
\end{equation}
is similar to the Nystr\"{o}m feature space (\ref{paperE:eqn:nystrom_feature_map}) with a different normalization scheme, as pointed out in section \ref{paperE:sec:model:speedup}.
The centralization of $\tilde\varphi(.)$ better conditions $\varphi(.)$ for a linear SVM solver, and normalization by the average $\ell_2$ norm is most useful for combining multiple similarity measures.
The BE-SVM classifier is
\begin{equation}f_\mathcal{B}(\mathbf{x}) = \mathbf{w}^\text{T} \varphi(\mathbf{x}) \geq 0\end{equation}
where $\mathbf{w}$ is solved by minimizing the primal BE-SVM objective
\begin{equation}
\frac{1}{2}\|\mathbf{w}\|_2^2 + C\sum_i \ell_H(y_i,f_\mathcal{B}(\mathbf{x}_i))^2
\label{paperE:besvm_primal}
\end{equation}
An $\ell_1$ regularizer results in sparser solutions, but with the cost of more expensive optimization than an $\ell_2$ regularization.
Therefore, for large scale scenarios, an $\ell_2$ regularization, combined with a reduced basis set $\mathcal{B}$, is preferred to an $\ell_1$ regularizer combined with a larger basis set.
Using multiple similarity measures is straightforward in BE-SVM.
The concatenated feature map $\varphi_M(\mathbf{x}) = \left[ \varphi^{(1)}(\mathbf{x})^\text{T}, \ldots, \varphi^{(M)}(\mathbf{x})^\text{T}\right]^\text{T}$ encodes the values of the $M$ similarity measures evaluated on the corresponding bases $\mathcal{B}^{(1)},\ldots,\mathcal{B}^{(M)}$.
In this work, we restrict the study to the case that the bases are shared among the $M$ similarity measures: \emph{i.e}\onedot} \def\Ie{\emph{I.e}\onedot $\mathcal{B}^{(1)}=\ldots=\mathcal{B}^{(M)}$.
\subsection{Indefinite Similarity Measures for Visual Recognition}
\label{paperE:sec:model:similarity_measures}
The lack of expressibility of the PSD kernels have been argued before \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot in \cite{sim_class,learning_kernels_from_indef,good_rec_nonmetric}.
For example, similarity measures which are not based on vectorial representations of data are most likely to be indefinite.
Particularly in computer vision, considering latent information results in lack of a fixed vectorial representation of instances, and therefore similarity measures based on latent information are most likely to be indefinite\footnote{Note that \cite{YangWVM12} and similar approaches use a PD kernel on fixed vectorial representation of the data, \emph{given the latent information}. The latent informations in turn are updated using an alterantive minimization approach. This makes the optimization non-convex, and differs from similarity measures which directly model latent informations.}.
A few applications of indefinite similarity measures in computer vision are pointed out below.
\cite{invariant_kernels} proposes (indefinite) jitter kernels for building desired invariances in classification problems.
\cite{AghazadehASC12} uses indefinite pairwise similarity measures with latent positions of objects for clustering.
\cite{deformable_models} considers deformation models for image matching.
\cite{graph_matching} defines an indefinite similarity measure based on explicit correspondences between pairs of images for image classification.
In this work, we consider similarity measures with latent deformations:
\begin{equation}
s(x_i,x_j) = \underset{z_i \in \mathcal{Z}(x_i),\,z_j \in \mathcal{Z}(x_j)}{\max} K_I(\mathbf{\phi}(x_i,z_i), \mathbf{\phi}(x_j,z_j)) +\mathcal{R}(z_i) + \mathcal{R}(z_j)
\label{paperE:eqn:indef_sim}
\end{equation}
where $K_I(.,.)$ is a similarity measure (potentially a PD kernel), $\mathbf{\phi}(x,z)$ is a representation of $x$ given the latent variable $z$, $\mathcal{R}(z)$ is a regularization term on the latent variable $z$, and $\mathcal{Z}(x)$ is the set of possible latent variables associated with $x$.
Specifically, when $\mathcal{R}(.) = 0$ and $\mathcal{Z}(x)$ involves latent positions, the similarity measure becomes similar to that of \cite{AghazadehASC12}.
When $\mathcal{R}(.) = 0$ and $\mathcal{Z}(x)$ involves latent positions and local deformations, it becomes similar to the zero order model of \cite{deformable_models}.
Finally, an MRF prior in combination with latent positions and local deformations gives a similarity measure, similar to that of \cite{graph_matching}.
The proposed similarity measure (\ref{paperE:eqn:indef_sim}) picks the latent variables which have the maximal (regularized) similarity values $K_I(.,.)$s.
This is in contrast to \cite{invariant_kernels} where the latent variables were suggested to be those which minimize a metric distance based on the kernel $K_I(.,.)$.
The advantage of a metric based latent variable selection is not so clear, while some works argue against unnecessary restrictions to metrics \cite{good_rec_nonmetric}.
Also, if $K_I(.,.)$ is not PSD, deriving a metric from it is at best expensive.
Therefore, the latent variables in (\ref{paperE:eqn:indef_sim}) are selected according to the similarity values instead of metric distances.
\subsection{Multi Class Classification}
\label{paperE:sec:model:multi_class}
SVMs are mostly known as binary classifiers.
Two popular extensions to the multi-class problems are one-v-res (1vR) and one-v-one (1v1).
The two simple extentions have been argued to perform as well as more sophisticated formulations \cite{defense_onevall}.
In particular, \cite{defense_onevall} concludes that in case of kernelized SVMs, in terms of accuracy they are both competitive, while in terms of training and testing complexities 1v1 is superior.
Therefore, we only consider 1v1 approach for kernelized SVM.
In case of linear SVMs however, 1v1 results in unnecessary overhead and 1vR is the algorithm of choice.
A 1vR BE-SVM can be expected to be both faster and to generalize better than a 1v1 BE-SVM where bases from all classes are used in each of the binary classifiers.
In case of 1v1 BE-SVM where only bases from the two classes under consideration are used in each binary classifier, there will be a clear advantage in terms of training complexity.
However, due to the reduction in the size of the basis set, the algorithm generalizes less in comparison to a 1vR approach.
Therefore, we only consider 1vR formulation for BE-SVM.
Table \ref{paperE:tab:complexity} summarizes the memory and computational complexity analysis for 1v1 kernelized SVM and 1vR BE-SVM.
Shown are the upper bounds complexities where we have considered $n$ to be the upper bound on $n_{sv}$.
\begin{table}
\centering
\begin{tabular}{|c||c|c|c|c|}
\hline
& \multicolumn{2}{|c|}{Training} & \multicolumn{2}{|c|}{Testing (per sample)} \\ \hline
& Memory & Computation & Memory & Computation \\ \hline
K SVM & $nM\bar{d}_\mathbf{\phi} + \frac{n^2}{C}$& $n^2M\bar{d}_K + \frac{n^3}{C}$ & $nM\bar{d}_\mathbf{\phi}$ & $nCM\bar{d}_K$\\ \hline
BE-SVM & $nM\bar{d}_\mathbf{\phi}+ nM|\mathcal{B}|$ & $nC|\mathcal{B}|M\bar{d}_K$ & $|\mathcal{B}|M\bar{d}_\mathbf{\phi}$ & $|\mathcal{B}|CM\bar{d}_K$ \\ \hline
\end{tabular}
\caption{Complexity Analysis for kernelized SVM and BE-SVM. The number of samples for each of the $C$ classes was assumed to be equal to $\frac{n}{C}$. $M$ is the number of kernels/similarity measures, $M\bar{d}_\mathbf{\phi}$ is the dimensionality of representations required for evaluating $M$ kernels/similarity measures, and $M\bar{d}_K$ is the cost of evaluating all $M$ kernels/similarity measures.}
\label{paperE:tab:complexity}
\end{table}
\subsection{Margin Analysis of Basis Expanding SVM}
\label{paperE:SM:sec:margin}
Both kernelized SVM and BE-SVM are max margin classifiers in their feature spaces.
The feature space of kernelized SVM $\psi_\mathcal{H}(.)$ is implicitly defined via the kernel function $k(.,.)$ while the feature space of the BE-SVM is explicitly defined via empirical kernel maps.
In order to derive the margin as a function of the data, we first need to derive the dual BE-SVM objective, where we assume a non-squared Hinge loss and unnormalized feature mappings $\tilde\varphi(.)$.
Borrowing from the representer theorem and considering the KKT conditions of the primal,
one can derive $\mathbf{w} = \sum_i y_i \beta_i \mathbf{\tilde\varphi}(\mathbf{x}_i)$, and consequently derive the BE-SVM dual objective which is similar to the dual SVM objective but with $\mathbf{K}_{ij} = \mathbf{\tilde\varphi}(\mathbf{x}_i)^\text{T} \mathbf{\tilde\varphi}(\mathbf{x}_j)$.
Let $\mathbf{S}_{BX}$ refer to the similarity of the data to the bases.
We can see that the margin of the BE-SVM, given the optimal dual variables $0\leq\mathbf{\beta}_i\leq C$, is $\left(\mathbf{\beta}^\text{T}\mathbf{Y}\mathbf{S}_{BX}^\text{T}\mathbf{S}_{BX}\mathbf{Y}\mathbf{\beta}\right)^{-1}$, as opposed to $\left(\mathbf{\alpha}^\text{T}\mathbf{Y}\mathbf{K}\mathbf{Y}\mathbf{\alpha}\right)^{-1}$ for the kernelized SVM, given the optimal dual variables $0\leq\mathbf{\alpha}_i\leq C$.
Furthermore, $\mathbf{S}_{BX}^\text{T}\mathbf{S}_{BX}$ is PSD, and that is BE-SVM's workaround for using indefinite similarity measures.
We analyze the margin of BE-SVM in case of unnormalized features ($\tilde\varphi(.)$ instead of $\varphi(.)$) and a non-squared Hinge loss.
Given the corresponding dual variables, the margin of the BE-SVM was mentioned to be
\begin{equation}
M_{BE}(\beta) = \left(\mathbf{\beta}^\text{T}\mathbf{Y}\mathbf{S}_{BX}^\text{T}\mathbf{S}_{BX}\mathbf{Y}\mathbf{\beta}\right)^{-1}
\end{equation}
as opposed to that of the kernelized SVM
\begin{equation}
M_{K}(\alpha) = \left(\mathbf{\alpha}^\text{T}\mathbf{Y}\mathbf{K}\mathbf{Y}\mathbf{\alpha}\right)^{-1}
\end{equation}
For comparison, the margin of the Nystr\"{o}mized method is
\begin{equation}
M_{N}(\alpha) = \left(\mathbf{\alpha}^\text{T}\mathbf{Y}\mathbf{K}_{XB}\mathbf{K}_{BB}^{-1}\mathbf{K}_{BX}\mathbf{Y}\mathbf{\alpha}\right)^{-1}
\end{equation}
\noindent\textbf{BE-SVM vs Kernelized SVM}:
When $s(.,.) = k(.,.)$ and all training exemplars are used as bases, the margin of the BE-SVM will be $\left(\mathbf{\beta}^\text{T}\mathbf{Y}\mathbf{K}^2\mathbf{Y}\mathbf{\beta}\right)^{-1}$.
Comparing to the margin of SVM, for the same parameter $C$ and the same kernel, it can be said that the solution (and thus the margin) of BE-SVM is even more derived by large eigenpairs, and even less by small ones.
It is straightforward to verify $\mathbf{K}^2=\sum_i \lambda_i^2 \mathbf{v}_i \mathbf{v}_i^\text{T}$.
Therefore, the contribution of large eigenpairs, that are $\{(\lambda_i,\mathbf{v}_i)|\lambda_i>1\}$, to $\mathbf{K}^2$ is amplified.
Similarly, the contribution of small eigenpairs, that are those with $\lambda_i<1$, to $\mathbf{K}^2$ is dampened.
\noindent\textbf{BE-SVM vs Nystr\"{o}mized method}:
When $s(.,.) = k(.,.)$ and a subset of training exemplars are used as bases (reduced settings), the resulting margin of BE-SVM is $\left(\mathbf{\beta}^\text{T}\mathbf{Y}\mathbf{K}_{XB}\mathbf{K}_{BX}\mathbf{Y}\mathbf{\beta}\right)^{-1}$.
Comparing to the margin of the Nystr\"{o}mized method, we can say that the most of the difference between the Nystr\"{o}mized method and BE-SVM, is the normalization by $\mathbf{K}_{BB}^{-1}$.
For covariance kernels, that the Nystr\"{o}mized method is most suitable for, $\mathbf{K}_{BB}$ is the covariance of the basis set in the feature space.
Therefore, it can be said that the normalization by $\mathbf{K}_{BB}^{-1}$ essentially de-correlates the bases in the feature space.
Although this is an appealing property, there is no associating RKHS with indefinite similarity measures and the de-correlation in such cases is non-trivial.
In case of covariance kernels, it can be said that BE-SVM assumes un-correlated bases, while bases are always correlated in the feature space.
As larger sets of bases usually result in more (non-diagonal) covariances, the un-correlated assumption is more violated with large set of bases.
The consequence is that in such cases, that are covariance kernels with large set of bases, BE-SVM can be expected to perform worse than the Nystr\"{o}mized method.
However, for sufficiently small set of bases, or in case of indefinite similarity measures, there is no reason for superiority of the Nystr\"{o}mized method.
In such cases and in practice, BE-SVM is competitive or better than the Nystr\"{o}mized method.
\subsubsection{Demonstration on 2D Toy data}
\label{paperE:SM:sec:margin:2D}
Figure \ref{paperE:SM:fig:besvm_ksvm_2D} visualizes the use of multiple Gaussian RBF kernels in BE-SVM and kernelized SVM.
We point out the following observations.
\begin{figure}
\centering
\begin{subfigure}[b]{0.32\textwidth}
\includegraphics[width=\textwidth]{Figures/KernelSVMgamma10100.pdf}
\caption{Kernelized SVM}
\label{paperE:SM:fig:besvm_ksvm_2D:ksvm}
\end{subfigure}
\begin{subfigure}[b]{0.32\textwidth}
\includegraphics[width=\textwidth]{Figures/BESVMDB651gamma10100.pdf}
\caption{BE-SVM dual objective}
\label{paperE:SM:fig:besvm_ksvm_2D:dual}
\end{subfigure}
\begin{subfigure}[b]{0.32\textwidth}
\includegraphics[width=\textwidth]{Figures/BESVMPB65gamma10100.pdf}
\caption{BE-SVM primal objective}
\label{paperE:SM:fig:besvm_ksvm_2D:primal_reduced}
\end{subfigure}
\caption{Demonstration of kernelized SVM and BE-SVM using two Gaussian RBF kernels with $\gamma_1=10,\gamma_2=10^2$ and $C=10$.
\ref{paperE:SM:fig:besvm_ksvm_2D:ksvm} is based on equally weighted kernels.
\ref{paperE:SM:fig:besvm_ksvm_2D:dual} is without normalization.
\ref{paperE:SM:fig:besvm_ksvm_2D:primal_reduced} is with normalization on 10\% of the data randomly selected as bases.
10 fold cross validation accuracy and the number of support vectors are averaged over $i=1:20$ scenarios based on the same problem but with different spatial noises.
The noise model for $i^\text{th}$ scenario is a zero mean Gaussian with $\sigma_i=10^{-2}i$.
The visualization is on the noiseless data for clarity.
Best viewed electronically.}
\label{paperE:SM:fig:besvm_ksvm_2D}
\end{figure}
1) the dual objective of BE-SVM (exact) tends to result in sparser solutions as measured by non-zero support vector coefficient (compare \ref{paperE:SM:fig:besvm_ksvm_2D:ksvm} with \ref{paperE:SM:fig:besvm_ksvm_2D:dual}).
We believe the main reason for this to be the modification of the eigenvalues as described in section \ref{paperE:SM:sec:margin}.
Note however that in order to classify a new sample, its similarity to all training data needs to be evaluated, irrespective of the sparsity of the BE-SVM solution (see equation (\ref{paperE:eqn:besvm_mult_kernel})).
In this sense, the BE-SVM dual objective results in completely dense solutions, similar to the primal BE-SVM objective without any basis reduction.
However, the solution can be made sparse by construction, by reducing the basis set, similar to the case with the primal BE-SVM objective.
We do not demonstrate this here, mainly because our main focus is on the (approximate) primal objective.
2) due to the definition of the (linear) kernel in BE-SVM (see equation (\ref{paperE:eqn:besvm_mult_kernel})), the solution of the BE-SVM has an inherent bias with respect to the (marginal) distribution of class labels.
In other words, the contribution of each class to the norm of $\tilde\varphi(.)$, and consequently to the value of $\tilde{K}(.,.)$, directly depends on the number of bases from each class.
Consequently, the decision boundary of BE-SVM is shifted towards the class with less bases: compare the decision boundaries on the left sides of \ref{paperE:SM:fig:besvm_ksvm_2D:ksvm} and \ref{paperE:SM:fig:besvm_ksvm_2D:dual}.
In experiments on CIFAR-10 dataset, as the number of exemplars from different classes are roughly equal, this did not play a crucial role.
\subsection{Related Work}
\label{paperE:sec:model:related_work}
There exists a body of work regarding the use of proximity data, similarity, or dissimilarity measures in classification problems.
\cite{similarity_svm} uses similarity to a fixed set of samples as features for a kernel SVM classifier.
\cite{pairwise_proximity} uses proximities to all the data as features for a linear SVM classifier.
\cite{classification_proximity_lp} uses proximities to all the data as features and proposes a linear program machine based on this representation.
In contrast, we use a normalization of the similarity of points to a subset of the data as features for a (fast, approximate) linear SVM classifier.
\section{Experiments}
\label{paperE:sec:experiments}
\subsection{Dataset and Experimental Setup}
\label{paperE:sec:experiments:dataset}
We present our experimental results on CIFAR-10 dataset \cite{Krizhevsky09learningmultiple}.
The dataset is comprised of 60,000 tiny $32\times32$ RGB images, 6,000 images for each of the 10 classes involved, divided into 6 folds with inequal distribution of class labels per fold.
The first 5 folds are used for training and the 6th fold is used for testing.
We use a modified version of the HOG feature \cite{DT05}, described in \cite{FelzenszwalbGMR10}.
For most of our experiments, we use HOG cell sizes of 8 and 4, which result in $31\times\frac{32}{8}^2=496$ and $31\times\frac{32}{4}^2=1984$ dimensional representation of each of the images.
Due to the normalization of each of the HOG cells, namely normalizing by gradient/contrast information of the neighboring cells, the HOG cells on border of images are not normalized properly.
We believe this to have a negative effect on the results, but as the aim of this paper is not to get the best results possible out of the model, we rely on the consistency of the normalization for all images to address this problem.
A possible fix is to up-sample images and ignore the HOG-cells at the boundaries, but we do not provide the results for such fixes.
For all the experiments, we center the HOG feature vector and scale feature vectors inversely by the average $\ell_2$ norm of the centered feature vectors, similar to the normalization of BE-SVM (\ref{paperE:eqn:besvm_feature_map_norm}).
This results in easier selection of parameters $C$ and $\gamma$ for SVM formulations.
Unless stated otherwise, we fix $C=2$ and $\gamma=1$ for kernelized SVM with Gaussian RBF kernels, and $C=1$ for the rest.
We use LibLinear \cite{REF08a} to optimize the primal linear SVM objectives with squared Hinge loss, similar to (\ref{paperE:besvm_primal}).
For kernelized SVM, we use LibSVM \cite{CC01a}.
We report multi-class classification results (0-1 loss) on the test set, where we used a 1 v 1 formulation for kernelized SVM, and 1 v all formulation for other methods.
\subsection{Baseline: SVM with Positive Definite Kernels}
Figure \ref{paperE:fig:multiple_sims_params_support:params} shows the performance of linear SVM (H4L and H8L) and kernelized SVM with Gaussian RBF kernel (K4R and K8R) as a function of number of parameters in the models.
The number of parameters for linear SVM is the input dimensionality, and for kernelized SVM it is the sum of $n_{sv} (d_\phi+1)$ where $d_\phi$ is the dimensionality of the feature vector the corresponding kernel operates on.
The 5 numbers for each model are the results of the model trained on $1,\ldots,5$ folds of the training data (each fold contains 10,000 samples).
Figure \ref{paperE:fig:multiple_sims_params_support:support} shows the performance kernelized SVM as a function of support vectors when trained on $1,\ldots,5$ folds.
Except the linear SVM with a HOG cell size of 8 pixels (496 dimensions) which saturates its performance at 4 folds, all models consistently benefit from more training data.
\subsection{BE-SVM with Invariant Similarity Measures}
\label{paperE:sec:experiments:invariant}
The general form of the invariant similarity measures we consider was given in (\ref{paperE:eqn:indef_sim}).
In particular, we consider rigid and deformable similarity measures where the smallest unit of deformation/translation is a HOG cell.
The rigid similarity measure models invariance to translations and is given by
\begin{equation}
K_R(x,y) = \max_{\mathbf{z}_R \in \mathcal{Z}_R} \sum_{\mathbf{c} \in \mathcal{C}} \phi_C(x,\mathbf{c})^T \phi_C(y,\mathbf{c} + \mathbf{z}_R)
\label{paperE:eqn:sim:rigid}
\end{equation}
where $\mathcal{Z}_R =\{(z_x,z_y)|z_x,z_y \in \{-h_R,\ldots,h_R\}\}$ allows a maximum of $h_R$ HOG cells displacements in $x,y$ directions, $\mathcal{C} = \{(x,y)|x,y \in \{h_1,\ldots,h_H\}$ is the set of indices of $h_H$ HOG cells in each direction, and $\phi_C(x,\mathbf{c})$ is the 31 dimensional HOG cell of $x$ located at position $\mathbf{c}$.
$\phi_C(x,\mathbf{c})$ is zero for cells outside $x$ (zero-padding).
$K_R(x,y)$ is the maximal cross correlation between $\phi(x)$ and $\phi(y)$.
The deformable similarity measure allows local deformations (displacements) of each of the HOG cells, in addition to invariance to rigid deformations
\begin{equation}
K_L(x,y) = \max_{z_R \in \mathcal{Z}_R} \sum_{\mathbf{c} \in \mathcal{C}} \max_{z_L \in \mathcal{Z}_L} \phi_C(x,\mathbf{c})^T \phi_C(y,\mathbf{c} + \mathbf{z}_R + \mathbf{z}_L)
\label{paperE:eqn:sim:nonrigid}
\end{equation}
where $\mathcal{Z}_L=\{(z_x,z_y)|z_x,z_y \in \{-h_L,\ldots,h_L\}\}$ allows a maximum of $h_L$ HOG cell local deformation for each of the HOG cells of $y$.
We consider a maximum deformation of 8 pixels \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot 2 HOG cells for a HOG cell size of 4 pixels.
Regularizing global or local deformations is straightforward in this formulation.
However, we did not notice significant improvements for the set of displacements we considered, which is probably related to the small size of the latent set suitable for small images in CIFAR-10.
Figure \ref{paperE:fig:full_basis_sel:full} shows the performance of BE-SVM using different similarity measures, when trained on the first fold.
It can be seen that the invariant similarity measures improve recognition performance.
Particularly, in absence of any other information, modelling rigid deformations (latent positions) seems to be much more beneficial than modelling local deformations.
An interesting observation is that aligning the data in higher resolutions is much more crucial: all models (linear SVM, kernelized SVM, and BE-SVM) suffer performace losses when the resolution is increased from a HOG cell size of 8 pixels to 4 pixels.
However, BE-SVM achieves significant performance gains by aligning the data in higher resolutions: compare H4L with H4(1,0) and H4(2,0), and H8L with H8(1,0).
\begin{figure}[t]
\centering
\begin{subfigure}[b]{0.49\textwidth}
\includegraphics[width=\textwidth]{Figures/BESVM_Full}
\caption{Full Bases}
\label{paperE:fig:full_basis_sel:full}
\end{subfigure}
\begin{subfigure}[b]{0.49\textwidth}
\includegraphics[width=\textwidth]{Figures/BasisSel_N100}
\caption{Basis Selection}
\label{paperE:fig:full_basis_sel:sel}
\end{subfigure}
\caption{Performance of BE-SVM as a function of different similarity measures when trained on the first fold. An H4 (H8) refers to a HOG cell size of 4 (8) pixels. L and R refer to linear and Gaussian RBF kernels respectively, and $(h_R,h_L)$ refers to a similarity measure with $h_R$ rigid and $h_L$ local deformations (\ref{paperE:eqn:sim:rigid}), (\ref{paperE:eqn:sim:nonrigid}).}
\label{paperE:fig:full_basis_sel}
\end{figure}
We tried training linear and kernelized SVM models by jittering the feature vectors, in the same manner that the invariant similarity measures do (\ref{paperE:eqn:sim:rigid}), (\ref{paperE:eqn:sim:nonrigid}); that is to jitter the HOG cells with zero-padding for cells outside images.
This resulted in significant performance losses for both linear SVM and kernelized SVM, while also siginificantly increasing memory requirement and computation times.
We believe the reason for this to be the boundary effects; which are also mentioned in previous work \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot \cite{invariant_kernels}.
We also believe that jittering the input images, in combination with some boundary heuristics (see section \ref{paperE:sec:experiments:dataset}), will improve the test performance (while significantly increasing training complexities), but we do not provide experimental results for such cases.
\subsection{Basis Selection}
\label{paperE:sec:experiments:basis_sel}
Figure \ref{paperE:fig:full_basis_sel:sel} shows accuracy of BE-SVM using different similarity measures and different basis selection strategies; for a basis size of $B=10\times100$ exemplars.
In the figure, `Rand' refers to a random selection of the bases, `Indx' refers to selection of samples according to their indices, `K KMed' refers to a kernel k-medoids approach based on the similarity measure, and `Nystrom' refers to selection of bases similar to the `Indx' approach, but with the Nystr\"{o}m normalization, using a spectrum clip fix for indefinite similarity measures(see section \ref{paperE:sec:model:speedup}).
The reported results for `Rand' method is averaged over 5 trials; the variance was not significant.
It can be observed that all methods except the `Nystrom' result in similar performances.
We also tried other sophisticated sample selection criteria, but observed similar behaviour.
We attribute this to little variation in the quality of exemplars in the CIFAR-10 dataset.
Having observed this, for the rest of sub-sampling strategies, we do not average over multiple random basis selection trials, but rather use the deterministic `Indx' approach.
The difference between normalization factors in BE-SVM and Nystr\"{o}m method (see section \ref{paperE:sec:model:speedup}) is evident in the figure.
The BE-SVM normalization tends to be consistently superior in case of indefinite similarity measures.
For PSD kernels (H4L, H8L, H4R, and H8R) , the Nystr\"{o}m normalization tends to be better in lower resolutions (H8) and worse in higher resolutions (H4).
We believe the main reason for this is to be lack of significant similarity of bases in higher resolutions in absence of any alignment.
In such cases, the low rank assumption of $\mathbf{K}$ \cite{Williams00theeffect} is violated, and normalization by a diagonally dominant $\mathbf{K}_{\mathbf{{m}}\mathbf{{m}}}$ will not capture any useful information.
In order to analyze how the performance of BE-SVM depends on the eigenvalues of the similarity measures, we provide the following eigenvalue analysis.
We compute the similarity of the bases to themselves -- corresponding to $\mathbf{K}_{\mathbf{{m}}\mathbf{{m}}}$ in (\ref{paperE:eqn:nystrom_feature_map}) -- and perform an eigen-decomposition of the resulting matrix.
Table \ref{paperE:tab:eigs} shows the ratio of negative eigenvalues: `NgRat'=$\frac{1}{B}\sum_i [\lambda_i<0]$ , and the relative energy of eigenvalues `NgEng'=$\frac{\sum_i |\lambda_i|[\lambda_i <0]}{\sum_i |\lambda_i|[\lambda_i >0]}$ as a function of various similarity measures for $B=10\times 100$ and a HOG cell size of 4.
The last two columns, namely `CorNyst' and `CorBE' reflect the correlation of the measured entities -- `NgRat' and `NgEng' -- to the observed performance of BE-SVM using the Nystr\"{o}m normalization and BE-SVM normalization.
We used Pearson's $r$ to measure the extent of linear dependence between the test performances and different normalization schemes.
It can be observed that: 1) both normalization schemes have a positive correlation with both the ratio of negative eigenvalues and their relative energy, and 2) BE-SVM normalization correlates more strongly with the observed entities.
From this, we conclude that negative eigenvectors contain discriminative information and that BE-SVM's normalization is more suitable for indefinite similarity measures.
We also experimented with spectrum flip and spectrum square methods for the Nystr\"{o}m normalization, but they generally provided slightly worse results in comparison to the spectrum clip technique.
\begin{table}
\centering
\begin{tabular}{|c|c|c|c|c|c|c||c|c|}
\hline
&{\small H4(0,0)}&{\small H4(0,1)}&{\small H4(1,0)}&{\small H4(1,1)}&{\small H4(2,0)}&{\small H4(2,1)}&{\small CorNyst}&{\small CorBE}\\ \hline
NgRat&.0&.26&.18&.25&.16&.30&.20 & .61\\ \hline
NgEng&.00&.04&.05&.05&.04&.07&.33 & .73\\ \hline
\end{tabular}
\caption{Eigenvalue analysis of various similarity measures based on HOG cell size 4.
}
\label{paperE:tab:eigs}
\end{table}
\subsection{Multiple Similarity Measures}
\label{paperE:section:multiple_sims}
Different similarity measures contain complementary information.
Fortunately, BE-SVM can make use of multiple similarity measures by construction.
To demonstrate this, using one fold of training data and $B=10\times 50$, we greedily -- in an incremental way -- augmented the similarity measures with the most contributing ones.
Using this approach, we found two (ordered) sets of similarity measures with complementary information: 1) a low-resolution set $\mathcal{M}_1=\{H8R, H8(1,0),H8(0,1)\}$, and 2) a two-resolution set $\mathcal{M}_2 = \{ H8R, H4(2,0), H4(0,1), H8(1,0)\}$.
Surprisingly, the two resolution sequence resembles those of the part based models \cite{FelzenszwalbGMR10}, and multi resolution rigid models \cite{AghazadehASC12} in that the information is processed at two levels: a coarser rigid `root' level and a finer scale deformable level.
We then trained BE-SVM models using these similarity measures for various sizes of the basis set, and for various sizes of training data.
Figures \ref{paperE:fig:multiple_sims_params_support:params} and \ref{paperE:fig:multiple_sims_params_support:support} show these results, where the BE-SVM models are trained on all 5 folds.
The shown number of supporting exemplars (and consequently the number of parameters) for BE-SVM are based on the size of the basis set.
It can be seen that using a basis size of $B=10\times 250$, the performance of the BE-SVM using more than 3 two-resolution similarity measures surpass that of the kernelized SVM trained on all the data and based on approximately $B=10\times4000$ support vectors.
Using low-resolution similarity measures, $B=10\times 500$ outperforms kernelized SVMs trained on up to 4 folds of the training data.
Furthermore, it can be observed that for the same model complexity, as measured either by the number of supporting exemplars, or by model parametrs, BE-SVM performs better than kernelized SVM.
\begin{figure}
\centering
\includegraphics[width=.66\linewidth]{Figures/BESVM_Mult12}
\caption{Performance of BE-SVM using multiple similarity measures for various sizes of the basis set.
Results with dotted, dashed, and solid lines represent 1, 3, and 5 folds worth of training data.
See text for analysis.
}
\label{paperE:SM:fig:multiple_sims}
\end{figure}
Figure \ref{paperE:SM:fig:multiple_sims} shows the performance of BE-SVM using different similarity measures for various basis sizes and for different training set sizes.
It can be observed that using (invariant) indefinite similarity measures can significantly increase the performance of the model: compare the red curve with any other curve with the same line style.
For example, using all the training data and a two resolution deformable approach results in 8-10\% improvements in accuracy in comparison to the best performing PSD kernel (H8R).
Furthermore, the two-resolution approach outperforms the single resolution approach by approximately 3-4\% accuracy (compare blue and black curves with the same line style).
Measured by model parameters, BE-SVM is roughly 8 times sparser than kernelized SVM for the same accuracy.
Measured by supporting exemplars, its sparsity increases roughly to 30.
We need to point out that different similarity measures have different complexities \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot H8(1,0) is more expensive to evaluate than K8R.
However, when the bases are shared for different similarity measures, CPU cache can be utilized much more efficiently as there will be less memory access and more (cached) computations.
\begin{figure}[t]
\centering
\begin{subfigure}[b]{0.49\textwidth}
\includegraphics[width=\textwidth]{Figures/BESVM_Params12}
\caption{Parameters vs Performance}
\label{paperE:fig:multiple_sims_params_support:params}
\end{subfigure}
\begin{subfigure}[b]{0.49\textwidth}
\includegraphics[width=\textwidth]{Figures/BESVM_Support12}
\caption{Supporting Exemplars vs Performance}
\label{paperE:fig:multiple_sims_params_support:support}
\end{subfigure}
\caption{Performance of BE-SVM vs model parameters for various sizes of the basis set, using multiple similarity measures.
Each curve for linear SVM (H4L, H8L) and kernelized SVM (K4R, K8R) represents the result for training on $1,\ldots,5$ folds of training data.
Each curve for BE-SVM shows the result for training model with a basis set of size $B=10\times\{25,50,100,250,500\}$ when trained on 5 folds of the training data.
}
\label{paperE:fig:multiple_sims_params_support}
\end{figure}
\subsubsection{Multiple Kernel Learning with PSD Kernels}
We tried Multiple Kernel Learning (MKL) for kernelized SVM with PSD kernels.
When compared to sophisticated MKL methods, we found the following procedure to give competitive performances, with much less training costs.
Defining $K_C(.,.) = \alpha K_1(.,.) + (1-\alpha) K_2(.,.)$, our MKL approach consists of performing a line search for an optimal alpha $\alpha\in\left\{0,.1,\ldots,1\right\}$ which results in best 5-fold cross validating performance.
Using this procedure, linear kernels were found not to contribute anything to Gaussian RBF kernels.
The optimal combination for high resolution and low resolution Gaussian RBF kernels (K4R and K8R) resulted in a performance gain of less than 0.5\% accuracy in comparison to K8R.
We founds this insignificant, and did not report its performance, considering the fact that the number of parameters increases approximately 4 times using this approach.
\subsection{BE-SVM's Normalizations}
It can be verified that in case of unnormalized features $\tilde\varphi^{(m)}(.)$, the corresponding Gram matrix will be
\begin{equation}
\begin{array}{ll}
\tilde K(\mathbf{x}_i,\mathbf{x}_j)
&= \sum_{m=1}^M\, \tilde K^{(m)}(\mathbf{x}_i,\mathbf{x}_j)
= \sum_{m=1}^M\, \sum_{b=1}^{B} \, s^{(m)}(\mathbf{b}_b,\mathbf{x}_i) s^{(m)}(\mathbf{b}_b,\mathbf{x}_j)
\end{array}
\label{paperE:eqn:besvm_mult_kernel}
\end{equation}
where $\tilde K^{(m)}$s are combined with equal weights, the value of each of which depends (locally) on how the similarities of $\mathbf{x}_i$ and $\mathbf{x}_j$ correlate with respect to the bases.
In the case of normalized features, the centered values of each similarity measure is weighted by $\left( \mathbb{E}_\mathcal{X}[\|\tilde\varphi - \mathbb{E}_\mathcal{X}[\tilde\varphi] \|] \right)^{-2}$ \emph{i.e}\onedot} \def\Ie{\emph{I.e}\onedot more global weight is put on (the centered values of) the similarity measures with smaller variances in similarity values.
While the BE-SVM's normalization of empirical kernel maps is not optimal for discrimination, it can be seen as a reasonable prior for combining different similarity measures.
Utilizing such a prior, in combination with linear classifiers and $\ell_P$ regularizers, has two important consequences: 1) the centering helps reduce the correlation between dimensions and the scaling helps balance the effect of regularization on different similarity measures, irrespective of their overall norms, and 2) such a scaling directly affects the parameter tuning for learning the linear classifiers: for all the similarity measures (and combinations of similarity measures) with various basis sizes, the same parameter: $C=1$ was used to train the classifiers.
While cross-validation will still be a better option, cross-validating for different parameters settings -- and specially when combining multiple similarity measures -- will be very expensive and prohibitive.
By using the BE-SVM's normalization, we essentially avoid searching for optimal combining weights for different similarity measures and also tuning for the $C$ parameter of the linear SVM training.
In this section, we quantitatively evaluate the normalization suggested for BE-SVM (\ref{paperE:eqn:besvm_feature_map_norm}), and compare it to a few other combinations.
Particularly, we consider various normalizations of the HOG feature vectors, and similarly, various normalization schemes for the empirical kernel map $\tilde{\varphi}$ (\ref{paperE:eqn:besvm_feature_map}).
We consider the following normalizations:
\begin{itemize}
\item No normalization (Unnorm)
\item Z-Scoring, namely centering and scaling each dimension by the inverse of its standard deviation (Z-Score)
\item BE-SVM normalization, namely centering and scaling all dimensions by the inverse average $\ell_2$ norm of the centered vectors (BE-SVM)
\end{itemize}
We report test performances for all combinations of normalizations for the feature vectors and the empirical kernel maps, for two cases: 1) when $C=1$, and 2) when the $C$ parameter is cross-validated from $\mathcal{C}=\{10^{-1}, 10^0, 10^1\}$.
In both cases, $|\mathcal{B}| = 10 \times 100$ bases were uniformly sub-sampled from the first fold of the training set (`Indx' basis selection).
\begin{figure}
\centering
\includegraphics[width=.49\linewidth]{Figures/Normalizations_C1}
\includegraphics[width=.49\linewidth]{Figures/Normalizations_CV}
\caption{Performance of BE-SVM for different normalization schemes of the feature vector and the empirical kernel map, and different similarity measures. ``F + K (P)'' in the legend reflects using F and K normalization schemes for the feature vectors and the empirical kernel maps respectively, which results in the average test performance of P (averaged over the similarity measures).}
\label{paperE:SM:fig:normalization_one_sim}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=.49\linewidth]{Figures/Normalizations_MS_C1}
\includegraphics[width=.49\linewidth]{Figures/Normalizations_MS_CV}
\caption{Performance of BE-SVM for different normalization schemes of the feature vector and the empirical kernel map, and different combinations of similarity measures. ``F + K (P)'' in the legend reflects using F and K normalization schemes for the feature vectors and the empirical kernel maps respectively, which results in the average test performance of P (averaged over the combinations of similarity measures).}
\label{paperE:SM:fig:normalization_multi_sim}
\end{figure}
Figure \ref{paperE:SM:fig:normalization_one_sim} shows the performance of BE-SVM in combination with different normalizations of the feature vectors and empirical kernel maps, and for different similarity measures.
On top, reported numbers are for $C=1$ while on the bottom, $C$ is cross validated.
It can be observed that the BE-SVM's normalization works best both for the feature and empirical kernel map normalizations.
Although z-scoring is more suitable for linear similarity measures (compare BE-SVM + BE-SVM with Z-SCORE + BE-SVM in H4L, H8L, H4(x,y) and H8(x,y)), overall BE-SVM's normalization of the feature space works better than the alternatives.
Particularly, in single similarity measure cases, it seems that normalizing the feature according to the BE-SVM's normalization is more important than normalizing the empirical kernel map.
While the cross-validation of the $C$ parameter marginally affects the performance, it does not change the conclusions drawn from the $C=1$ case.
Figure \ref{paperE:SM:fig:normalization_multi_sim} shows the performance of BE-SVM in combination with different normalizations of the feature vectors and empirical kernel maps, and for different combinations of similarity measures (the sequence of greedily augmented similarity measures $\mathcal{M}_2$: the set of two resolution similarity measures described in Section \ref{paperE:section:multiple_sims}).
It can be observed that BE-SVM's normalization of the kernel map is much more important and effective when combining multiple similarity measure (compare to Figure \ref{paperE:SM:fig:normalization_one_sim}) .
These observations quantitatively motivate the use of BE-SVM's normalization with the following benefits, at least on the dataset we experimented on:
\begin{itemize}
\item It removes the need for cross-validation for tuning the $C$ parameter, and mixing weights for different similarity measures.
\item As the feature vector is centered and properly scaled, the linear SVM solver converges much faster than the unnormalized case, or when $C>>1$.
\item It results in robust learning of BE-SVM which can efficiently combine different similarity measures \emph{i.e}\onedot} \def\Ie{\emph{I.e}\onedot RBF kernels (H8R), and linear deformable similarity measures (H4(2,0), H4(0,1), H8(1,0)).
\end{itemize}
\section{Conclusion}
\label{paperE:sec:conclusion}
We analyzed scalable approaches for using indefinite similarity measures in large margin scenarios.
We showed that our model based on an explicit basis expansion of the data according to arbitrary similarity measures can result in competitive recognition performances, while scaling better with respect to the size of the data.
The model named Basis Expanding SVM was thoroughly analyzed and extensively tested on CIFAR-10 dataset.
In this study, we did not explore basis selection strategies, mainly due to the small intra-class variation of the dataset.
We expect basis selection strategies to play a crucial role in the performance of the resulting model on more challenging datasets \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot Pascal VOC or ImageNet.
Therefore, an immediate future work is to apply BE-SVM to larger scale and more challenging problems \emph{e.g}\onedot} \def\Eg{\emph{E.g}\onedot object detection, in combination with data driven basis selection strategies.
|
\section{Introduction {and main results}}
\setcounter{equation}{0}
In this paper, we consider the problem:
\bea
&&u_t= \Delta u+f(x)(1-u)^{-p},\quad {x\in\Omega},\; t>0,\label{pdeGen}\\
&&u=0, \quad x\in\partial\Omega,\; t>0,\label{bcGen}\\
&&u(x,0)=0,\quad x\in\Omega,\label{icGen}
\eea
where $\Omega$ is a bounded domain of $\BR^n$ ($n\ge 1$), {of class $C^{2+\nu}$ for some $\nu>0$,}
\be\label{Hypfp}
\hbox{ $p>0$ and $f$ is a nonnegative, H\"older continuous function on~$\overline\Omega$, with $f\not\equiv 0$.}
\ee
A typical case of interest is the following
\bea
&&u_t={\Delta u}+\lambda {|x|^m}(1-u)^{-2},\quad {x\in\Omega}, \; t>0,\label{pde}\\
&&u=0, \quad {x\in\partial\Omega},\quad t>0,\label{bc}\\
&&u(x,0)=0,\quad {x\in\Omega},\label{ic}
\eea
where $m, \lambda$ are positive constants.
This problem arises in the study of modeling the dynamic deflection of an elastic membrane inside a
micro-electro mechanical system~(MEMS).
The full model is
\be\label{mod}
\epsilon u_{tt}+u_t=\Delta u+\frac{\lambda g(x)}{(1-u)^2\Big(1+\alpha\int_{\Omega}\frac{1}{1-u}dx\Big)^2},
\quad x\in \Omega, \; t>0,
\ee
where $\epsilon$ is the ratio of the interaction due to the inertial and damping terms,
$\lambda$ is {proportional} to the applied voltage, $u$ is the deflection of the membrane
{(the natural physical dimension being thus $n=2$).}
The function $g(x)$, {called the permittivity profile}, represents varying dielectric properties of the membrane.
One of physically suggested dielectric profiles is the power-law profile
$g(x)=|x|^m$ with $m>0$.
The integral in (\ref{mod}) arises due to the fact that the device is embedded
in an electrical circuit with a capacitor of fixed capacitance.
The parameter $\alpha$ denotes the ratio of this fixed capacitance to a reference capacitance of the device.
{As for the initial condition (\ref{icGen}), it means that the membrane has initially no deflection,
the voltage being switched on at $t=0$.}
For the details of background and derivation of this model, we refer the reader to \cite{PT01, JAP-DHB02, FMPS07}.
{The case when $\epsilon=0$ is studied in \cite{GHW08,GK,G-s} for $\alpha>0$ and $f$ a constant.
We shall here concentrate on the case when $\epsilon=\alpha=0$ (so that there is no capacitor in the circuit)
and $f$ is nonconstant.}
It has been studied extensively for past years, see, e.g., the works \cite{G1, YG-ZP-MJW06, GG08, G08, G08A, KMS08,YZ10}.
For the study of stationary solutions, we refer to \cite{L89, G97, EGG07,GG07,E08, EG08, GW08, KMS08,YZ10}.
\medskip
By the standard parabolic theory, there exists a unique classical solution of (\ref{pdeGen})-(\ref{icGen}) in a short time interval.
Also, by the strong maximum principle, we have $u>0$ in {$\Omega$ for $t>0$.}
Moreover, the solution $u$ of (\ref{pdeGen})-(\ref{icGen}) can be continued as long as $\max_{x\in\overline\Omega}u(x,t)<1$.
We shall let $[0,T)$ be the maximal existence time interval of $u$, where $T\le\infty$.
If $T<\infty$, then quenching occurs in finite time, i.e.
$$\limsup_{t\to T^-}\ \{\max_{x\in\overline\Omega}u(x,t)\}=1.$$
It is well known {(see, e.g., \cite{EGG,KMS08} and the references therein)}
that the solution of (\ref{pdeGen})-(\ref{icGen}) quenches in finite time when $\lambda$ is sufficiently large.
A point $x=x_0$ is a {\it quenching point} if there exists a sequence $\{(x_n,t_n)\}$
in $\overline\Omega\times(0,T)$ such that
$$\hbox{$x_n\to x_0$, $t_n\uparrow T$ and $u(x_n,t_n)\to 1$ as $n\to\infty$.}$$
The set of all quenching points is called the quenching set, denoted by $\mathcal{Q}$.
In the context of MEMS, quenching corresponds to a touchdown phenomenon.
\medskip
{Note that in the typical case of \eqref{pde}, there is no source at $x=0$ due to the spatially dependent coefficient $|x|^m$.}
A long-standing open problem, {even in one space dimension,} is to determine whether or not $x=0$ is a quenching point.
More generally, for problem (\ref{pdeGen})-(\ref{icGen}), the question is whether a point $x_0$ such that $f(x_0)=0$
can be a quenching point.
In \cite{GG08,G08}, under the assumption that the quenching set is a compact subset of~${\Omega}$,
it is shown that {$x_0$ is not a quenching point if $f(x_0)=0$.}
On the other hand, {\color{black} the compactness assumption was proved in \cite{G08} by adapting a moving plane argument from \cite{FM,G1}}
when $f$ is constant or, more generally, when $f$ is nonincreasing as one approaches the boundary.
However, for {the typical} problem (\ref{pde})-(\ref{ic}) it is unknown whether the quenching set is compact.
Actually, supported by numerical evidence provided in \cite{YG-ZP-MJW06}, the following conjecture was made
(see~\cite{GG08,G08, EGG}):
\medskip
{\bf Conjecture.} {\it The point $x=0$ is not a quenching point for problem (\ref{pde})-(\ref{ic}).}
\medskip
{\bf In the present paper, we give an affirmative answer to this conjecture,
as well as for the case of general $f$, in any space dimension.}
Our main result is the following.
\begin{theorem} \label{MainThm}
Assume (\ref{Hypfp}) and let the solution $u$ of problem (\ref{pdeGen})-(\ref{icGen})
be such that $T<\infty$.
If $x_0\in \Omega$ is such that $f(x_0)=0$, then $x_0$ is not a quenching point.
\end{theorem}
In particular, as a special case, we have that $0$ is not a quenching point for problem (\ref{pde})-(\ref{ic}).
Actually, we have been able to answer this question {\it independently} of the compactness issue of quenching set.
In fact, as a key-step -- of independent interest -- to the proof of Theorem~1,
we prove the following estimate, which in particular guarantees that
the quenching rate is of {\it type I} on any compact subset of $\Omega$.
{In what follows, we denote
$$\delta(x)={\rm dist}(x, \partial\Omega),\quad x\in\overline\Omega,$$
the function distance to the boundary.}
\begin{theorem} \label{MainProp}
Assume (\ref{Hypfp}) and let the solution $u$ of problem (\ref{pdeGen})-(\ref{icGen})
be such that $T<\infty$.
Then there exists a constant $\gamma>0$ {\color{black} (independent of $x,t$)} such that
\be\label{rate}
1-u(x,t)\ge \gamma\, \delta(x)\,(T-t)^{1/(p+1)},\quad x\in \Omega,\ 0<t<T.
\ee
\end{theorem}
{Theorem~\ref{MainProp} will be proved via a nontrivial modification of the Friedman-McLeod method (\cite{FM}, see also \cite{G1}).
Once Theorem~\ref{MainProp} is proved, Theorem~\ref{MainThm} will be deduced by constructing a suitable local supersolution.}
\medskip
The compactness of the quenching set remains an open question.
In particular, we do not know if Theorem~\ref{MainThm} remains true if $f(x_0)=0$ with $x_0\in\partial\Omega$
(in other words, can a zero {\it boundary} point of the permittivity profile
be a quenching point ?).
As mentioned above, this cannot occur if we assume in addition that $f$
is nonincreasing as one approaches the boundary.
{Actually, as a consequence of Theorem~\ref{MainProp} and of suitable comparison arguments, we
have been able to obtain two further criteria for the quenching set to be compact.
We note that, unlike in the aforementioned criterion, we do not require any convexity of the domain~$\Omega$.}
\begin{theorem} \label{BoundaryResult}
Assume (\ref{Hypfp}) and let the solution $u$ of problem (\ref{pdeGen})-(\ref{icGen})
be such that $T<\infty$. Assume either
\be\label{HypBdry1}
0<p<1,
\ee
or
\be\label{HypBdry}
f(x)=o\bigl(\delta^{p+1}(x)\bigr)\qh{ as } \delta(x)\to 0.
\ee
Then quenching does not occur near the boundary, i.e. $\mathcal{Q}\subset\Omega$.
\end{theorem}
{Going back to MEMS modeling, it seems that {\color{black} Theorem~\ref{MainThm} and the case of \eqref{HypBdry} in Theorem~\ref{BoundaryResult}}
may be of some importance in applications,
at least from the qualitative point of view (see \cite{YG-ZP-MJW06, EGG} for more details).
This is especially true for particular devices of MEMS type such as micro-valves, where the touchdown behavior is explicitly exploited,
since the touchdown or quenching set then corresponds to the lid or closing area of the valve.
As a consequence of our results, we see that the latter has to be part of the positive set of the function $f$.
The choice of $f$, through an appropriate repartition of the dielectric coating, can thus be used in the optimal design of the microvalve.
In this respect, it would be desirable to gain further information about the structure of the quenching set,
but this seems a difficult mathematical problem for nonconstant $f$, even in one space dimension.
}
\medskip
{{\bf Remark.}} We point out that Theorems~\ref{MainThm}--\ref{BoundaryResult} still hold if we replace {\color{black} as in \cite{KMS08}}
the zero initial data by a nonnegative
$C^2$ function $u_0$ such that $u_0<1$ in $\overline\Omega$, $\Delta u_0+f(x)(1-u_0)^{-p}\ge 0$ in $\Omega$
and $u_0=0$ on $\partial\Omega$.
{Indeed, this assumption guarantees that $u_t>0$ and the proofs can then be modified
in a straightforward way.}
\section{Proof of Theorem 2}\label{point}
\setcounter{equation}{0}
\subsection{General strategy and basic computation}
When the compactness of the quenching set is known, type I estimates can
be proved by means of the maximum principle applied, {in a strict subdomain of $\Omega$,}
to the well-known auxiliary function (cf. \cite{FM, G1}):
$$J(x,t):=u_t-\varepsilon(1-u)^{-p},$$
where $\varepsilon$ is a small positive constant.
In the present situation, the possible noncompactness of the quenching set prevents one to verify that
$J\ge 0$ on the boundary of any subdomain of $\Omega$
and the method is not directly applicable.
To overcome this, our basic idea is to consider a modified function $J$ as follows:
\be \label{defJah}
J{\color{black} (x,t)}=u_t-\varepsilon a(x)h(u),
\ee
where $a(x)$ is an auxiliary function such that $a=0$ on $\partial\Omega$, hence also $J=0$.
The construction is delicate and requires specific properties for $a$, which will be given later.
As for the function $h(u)$, it will be a perturbation of the nonlinearity,
namely
\be \label{defhJ}
h(u)=(1-u)^{-p}+
(1-u)^{-q},\quad 0<q<p.
\ee
Before specializing, we first present the basic computations.
\begin{lemma} \label{lem1}
Let $J, h$ be given by (\ref{defJah})-(\ref{defhJ}), where $a\in C^2(\Omega)$ is a nonnegative function.
Then
\be\label{identJR}
J_t- \Delta J -pf(x)(1-u)^{-p-1}J = \varepsilon R,
\ee
where
\be\label{defJR}
R=
(p-q)a(x)f(x)(1-u)^{-p-q-1}
+ah''(u)|\nabla u|^2
+2h'(u)\nabla a\cdot\nabla u+h(u)\Delta a.
\ee
Moreover, {$h''>0$ and,} at any point $x\in\Omega$ such that $a(x)>0$, we have
\be\label{estimJR}
R\ge
\underbrace{(p-q)a(x)f(x)(1-u)^{-p-q-1}}_{{\mathcal{T}}_1}
+\underbrace{h(u)\Delta a}_{{\mathcal{T}}_2}
-\underbrace{\frac{{h'}^2(u)|\nabla a|^2}{ah''(u)}}_{{\mathcal{T}}_3}.
\ee
\end{lemma}
\begin{proof}
We compute
\beaa
J_t&=&u_{tt}-\varepsilon a(x)h'(u) u_t \\
\nabla J&=&\nabla u_t-\varepsilon \bigl(a(x)h'(u)\nabla u+h(u)\nabla a(x)\bigr)\\
\Delta J &=&\Delta u_t-\varepsilon \bigl(a(x)h'(u)\Delta u+a(x)h''(u)|\nabla u|^2+2h'(u)\nabla a(x)\cdot\nabla u+h(u)\Delta a(x)\bigr).
\eeaa
Setting $g(u)=(1-u)^{-p}$ and omitting the variables $x, u$ without risk of confusion, we get
\beaa
J_t-\Delta J
&=&(u_t-\Delta u)_t-\varepsilon ah' (u_t-\Delta u)+\varepsilon (ah'' |\nabla u|^2+2h'\nabla a\cdot\nabla u+h \Delta a), \\
\noalign{\vskip 1mm}
&=&f(x) g' u_t-\varepsilon f(x) ah' g+\varepsilon (ah'' |\nabla u|^2+2h'\nabla a\cdot\nabla u+h \Delta a).
\eeaa
Using $u_t=J+\varepsilon ah$, we have
$$J_t- \Delta J -f(x) g' J
= \varepsilon R,$$
where
$$R=f(x)a (g' h-h' g)+ ah'' |\nabla u|^2+2h'\nabla a\cdot\nabla u+h \Delta a.$$
On the other hand, we have
$$
\begin{array}{lll}
g'h-h' g
&=p(1-u)^{-p-1}\bigl((1-u)^{-p}+
(1-u)^{-q}\bigr)\cr
\noalign{\vskip 1mm}
&\ \ \ -(1-u)^{-p}\bigl(p(1-u)^{-p-1}+
q(1-u)^{-q-1}\bigr) \cr
\noalign{\vskip 1mm}
&=
(p-q)(1-u)^{-p-q-1},
\end{array}
$$
hence (\ref{defJR}).
Finally, {since $h''>0$,} for all $x\in \Omega$ such that $a(x)>0$, we may write
$$R=
(p-q)af(x)(1-u)^{-p-q-1} + h\Delta a+ah''\Bigl[|\nabla u|^2+2\frac{h'\nabla a\cdot\nabla u}{ah''}\Bigr],$$
hence (\ref{estimJR}).
\end{proof}
\subsection{Construction of the function $a(x)$}
We see that, in order to guarantee $R\ge 0$,
{the (negative) term ${\mathcal{T}}_3$ on the RHS of (\ref{estimJR}) must be absorbed by
a positive contribution coming either:
\smallskip
$\bullet$ from the term ${\mathcal{T}}_1$ (generated by the perturbation in (\ref{defhJ})), provided $f(x)>0$; or
\smallskip
$\bullet$ from the term ${\mathcal{T}}_2$, provided $\Delta a(x)>0$.}
\smallskip
Since $a(x)$ is nonnegative and vanishes at the boundary, we cannot have $\Delta a>0$ everywhere.
Actually, we shall consider functions $a(x)$ which are positive in $\Omega$ and suitably convex everywhere except on a small ball $B$,
where $f$ is uniformly positive.
Also, it will be necessary to split the parabolic cylinder
$\Omega\times (T/2,T)$ into suitable sub-regions,
taking into account the ``large'' and ``small'' parts of
the function $u(x,t)$.
\medskip
The following essential lemma gives the construction of the appropriate function~$a(x)$.
\begin{lemma} \label{lem2}
Let $h$ be given by (\ref{defhJ}).
Let $x_0\in\Omega, \rho>0$ with $\overline B(x_0,\rho)\subset \Omega$,
and denote the open set
$$\Omega_{x_0,\rho}=\Omega\setminus \overline B(x_0,\rho).$$
Then there exists a function $a\in C^1(\overline\Omega)\cap C^2(\Omega)$ with the following properties:
\be\label{prop-ah1}
hh''a\Delta a-{h'}^2|\nabla a|^2\ge 0 \qh{for all $x\in\overline{\Omega_{x_0,\rho}}$ and all $0\le u<1$,}
\ee
\be\label{prop-ah2}
C_1\delta^{p+1}(x)\le a(x)\le C_{2}\delta^{p+1}(x) \qh{ for all $x\in\overline\Omega$},
\ee
for some constants $C_1, C_2>0$.
\end{lemma}
\begin{proof}
For $0<u<1$, computing
\bea
h'(u)&=&p(1-u)^{-p-1}+
q(1-u)^{-q-1}, \label{comput-hprime1} \\
\noalign{\vskip 1mm}
h''(u)&=&p(p+1)(1-u)^{-p-2}+
q(q+1)(1-u)^{-q-2}, \label{comput-hprime2}
\eea
it follows that
\beaa
hh''
&=&p(p+1)(1-u)^{-2p-2}+
(p(p+1)+q(q+1))(1-u)^{-p-q-2}\cr
\noalign{\vskip 2mm}
&&\ +
q(q+1)(1-u)^{-2q-2}\cr
\noalign{\vskip 2mm}
&\ge&
\frac{p+1}{p}\Bigl[p^2(1-u)^{-2p-2}+2
pq(1-u)^{-p-q-2}+
q^2(1-u)^{-2q-2}\Bigr],
\eeaa
{where we used} $p(p+1)+q(q+1)> 2(p+1)q$ due to $0<q<p$. Therefore, we have
\be\label{comp-h}
hh''\ge \frac{p+1}{p}(h')^2 \qh{ for $0\le u<1$.}
\ee
Next we introduce a suitable harmonic function $\phi$, namely, the unique solution of the problem
\beaa
&&\Delta \phi=0,\quad x\in\Omega_{x_0,\rho}, \\
&&\phi=0, \quad x\in \partial\Omega, \\
&&\phi=1,\quad x\in \partial B(x_0,\rho).
\eeaa
The function $\phi$ is smooth and, by the strong maximum principle and the Hopf Lemma,
we have $0<\phi<1$ in $\Omega_{x_0,\rho}$ {\color{black} and
\be\label{prop-ah2b}
c_1\delta(x)\le \phi(x)\le c_2\delta(x),\quad x\in \overline{\Omega_{x_0,\rho}}
\ee
for some positive constants $c_1,c_2$.}
Then we set
\be\label{defaphi}
a(x)=\phi^{p+1}(x),\quad x\in\overline{\Omega_{x_0,\rho}}.
\ee
Since $a\in C^2(\Omega_{x_0,\rho})$, the boundary $\partial B(x_0,\rho)$ is smooth and $a=1$ on $\partial B(x_0,\rho)$,
the function $a$ can be extended in $B(x_0,\rho)$
in such a way that $a\in C^2(\Omega)$ and $a>0$ in $\Omega$.
On the other hand, on $\Omega_{x_0,\rho}$, we compute:
$$\nabla a=(p+1)\phi^p\nabla\phi,\qquad
\Delta a=(p+1)[\phi^p\Delta\phi+p\phi^{p-1}|\nabla\phi|^2]=p(p+1)\phi^{p-1}|\nabla\phi|^2,$$
{\color{black} hence}
\be\label{comp-a}
a\Delta a=\frac{p}{p+1} |\nabla a|^2 \qh{on $\Omega_{x_0,\rho}$.}
\ee
Combining (\ref{comp-h}) and (\ref{comp-a}), we get (\ref{prop-ah1}).
Property (\ref{prop-ah2}) follows from (\ref{prop-ah2b}), (\ref{defaphi})
{and $a>0$ in $\Omega$.}
\end{proof}
Before going further, let us recall the following useful lower bound on $u_t$.
\begin{lemma} \label{utHopf}
There exists a constant $c_0>0$ such that
\be\label{lowerut}
u_t\ge c_0 \delta(x) \qh{ on $\Omega\times [T/2,T)$.}
\ee
\end{lemma}
\begin{proof}
{\color{black} Although the proof of the lemma is standard (cf. \cite{FM,G1,G08}), we provide a proof here for completeness.}
Setting $v=u_t$, we see that $v$ satisfies
\beaa
&&v_t={\Delta u}+pf(x)(1-u)^{-p-1}v,\quad x\in\Omega,\; 0<t<T,\\
&&v=0,\quad x\in\partial\Omega,\ 0<t<T,\\
&&v(x,0)=f(x),\quad x\in\Omega,
\eeaa
so that $v=u_t\ge 0$ in $Q_T:=\Omega\times(0,T)$ by the maximum principle.
Applying the maximum principle again, we deduce that $u_t\ge z$ in $Q_T$,
where $z$ is the solution of the heat equation in $Q_T$, with zero boundary condition and
initial condition $z(\cdot,0)=f$. Since $z$ satisfies the estimate~(\ref{lowerut})
in virtue of the Hopf lemma and the strong maximum principle, so does~$u_t$.
\end{proof}
\subsection{Proof of Theorem 2}
$ $
\medskip
{\bf Step 1.} {\it Preparations.}
Since $f\ge 0$ and $f\not\equiv 0$ is continuous,
we may pick a point $x_0\in \Omega$ and
$\rho>0$ such that $B(x_0,2\rho)\subset\Omega$ and
\be\label{inffx0
\sigma_1:
\inf_{x\in \overline B(x_0,\rho)}f(x)>0.
\ee
We then take $a\in C^1(\overline\Omega)\cap C^2(\Omega)$ as given by Lemma~\ref{lem2},
and define $J$ by
$$J(x,t)=u_t-\varepsilon a(x)h(u),$$
with $\varepsilon>0$ to be fixed later and
\be\label{defh2}
h(u)=(1-u)^{-p}+(1-u)^{-q}, \quad \hbox{\color{black} $0\le u<1$,\quad where $q=p/2$}
\ee
{\color{black} (any choice of $q\in (0,p)$ would do).}
Note that
\be\label{infax0}
\sigma_2:=
\inf_{x\in \overline B(x_0,\rho)}a(x)>0.
\ee
Next, we split the cylinder $\Sigma:=\Omega\times (T/2,T)$ into three
sub-regions as follows:
$$\Sigma_1=\bigl(\Omega\setminus \overline B(x_0,\rho)\bigr)\times (T/2,T),$$
$$\Sigma_2^\eta=\bigl\{(x,t)\in \overline B(x_0,\rho)\times(T/2,T);\ u(x,t)\ge 1-\eta \bigr\}$$
and
$$\Sigma_3^\eta=\bigl\{(x,t)\in \overline B(x_0,\rho)\times(T/2,T);\ u(x,t)< 1-\eta \bigr\},$$
where the number $\eta\in (0,1)$ will be specified later on.
\medskip
{\bf Step 2.} {\it Parabolic inequality for $J$ in the sub-regions
$\Sigma_1$ and $\Sigma_2^\eta$.}
It follows from properties~(\ref{estimJR}) in Lemma~\ref{lem1} and (\ref{prop-ah1}) in Lemma~\ref{lem2}, along with $a>0$, $f\ge 0$ {\color{black} in $\Omega$ and} $h''>0$, that
\be\label{ineqJSigma1}
J_t- \Delta J -pf(x)(1-u)^{-p-1}J\ge 0 \qh{ in $\Sigma_1$.}
\ee
Next, in view of {\color{black} \eqref{defh2} and} (\ref{comput-hprime1}),
we have
\beaa
|h\Delta a|\le {\color{black} C_3}(1-u)^{-p},\quad |h'\nabla a|\le {\color{black} C_3}(1-u)^{-p-1}
\qh{{in $\Sigma$}}
\eeaa
{\color{black} for some positive constant $C_3$ independent of $\varepsilon,\eta$.
Also, from (\ref{comput-hprime2}) and \eqref{infax0} we get
\beaa
a h''\ge \sigma_2 p(p+1)(1-u)^{-p-2} \qh{ in $\overline{B}(x_0,\rho)\times(0,T)$.}
\eeaa
}
Consequently, recalling the definition of $R$ in Lemma~\ref{lem1},
it follows from (\ref{estimJR}), (\ref{inffx0}) and (\ref{infax0}) that
\beaa
(1-u)^{p+q+1}R
&\ge& (p-q)f(x)a +h\Delta a(1-u)^{p+q+1}-\frac{(h'|\nabla a|)^2}{ah''}(1-u)^{p+q+1}\cr
\noalign{\vskip 1mm}
&\ge& (p-q)\sigma_1\sigma_2
-{\color{black} C_4} (1-u)^{q+1}
\ge (p-q)\sigma_1\sigma_2
-{\color{black} C_4}\eta^{q+1} \qh{ in $\Sigma_2^\eta$,}
\eeaa
{\color{black} for some positive constant $C_4$ independent of $\varepsilon,\eta$.}
Owing to (\ref{identJR}), we may thus choose $\eta\in (0,1)$ small,
independent of $\varepsilon$, such that
\be\label{ineqJSigma2}
J_t- \Delta J -pf(x)(1-u)^{-p-1}J\ge 0
\qh{ in $\Sigma_2^\eta$.}
\ee
\medskip
{\bf Step 3.} {\it Control of $J$ on $\Sigma_3^\eta$ and conclusion.}
Now that $\eta$ has been fixed, using (\ref{prop-ah2}),
(\ref{lowerut}) and {\color{black} \eqref{defh2}},
we may choose $\varepsilon>0$ small enough, so that
\be\label{ineqJSigma3}
J\ge \delta(x)\Bigl[c_0 - 2C_2\varepsilon\delta^p(x) (1-u)^{-p}\Bigr]
\ge \delta(x)\Bigl[c_0 - 2C_2\varepsilon\delta^p(x) \eta^{-p}\Bigr]\ge 0
\qh{in $\Sigma_3^\eta$}
\ee
and
\be\label{ineqJinit}
J(x,T/2)\ge \delta(x)\Bigl[c_0 -2
C_2\varepsilon\delta^p(x) \bigl(1-\|u(\cdot,T/2)\|_\infty\bigr)^{-p}\Bigr]\ge 0
\qh{in $\overline\Omega$,}
\ee
{\color{black} where $c_0$ is the constant in Lemma~\ref{utHopf} and $C_1,C_2$ are the constants in \eqref{prop-ah2}.}
Observe that, as a consequence of (\ref{ineqJSigma3})
and $\Sigma=\Sigma_1\, \cup \,\Sigma_2^\eta\,\cup\, \Sigma_3^\eta$,
we have
\be\label{Csq-ineqJSigma}
\bigl\{(x,t)\in \Sigma;\ J(x,t)<0\bigr\}\subset \Sigma_1\cup \Sigma_2^\eta.
\ee
Also, since $a=0$ on $\partial\Omega$, we have
\be\label{ineqJboundary}
J=0 \qh{ on $\partial\Omega\times (T/2,T)$.}
\ee
On the other hand, by standard parabolic regularity,
we observe that
$$J\in C^{2,1}(\Sigma)\cap C(\overline\Omega\times [T/2,T)).$$
It follows from (\ref{ineqJSigma1})-(\ref{ineqJSigma2}),
(\ref{ineqJinit})-(\ref{ineqJboundary}) and
the maximum principle
(see e.g. \cite[Proposition 52.4 and Remark 52.11(a)]{QS})
that
$$J\ge 0 \qh{ in $\Sigma$.}$$
Then, for $T/2<t<s<T$ and $x\in \Omega$, taking (\ref{prop-ah2}) into account, an integration in time gives
\beaa
(1-u(x,t))^{p+1}
&\ge& (1-u(x,s))^{p+1}+C_1\varepsilon(p+1)\delta^{p+1}(x)(s-t) \\
&\ge& C_1\varepsilon(p+1)\delta^{p+1}(x)(s-t).
\eeaa
Letting $s\to T$, we finally deduce (\ref{rate})
in $\Sigma$, hence in $\Omega\times (0,T)$.
{\color{black} Note that the constants $C_1,\varepsilon$ are independent of $x,t$, so is the constant $\gamma$ in \eqref{rate}.}
\qed
\section{Proof of Theorem 1}
\setcounter{equation}{0}
With the type~I estimate \eqref{rate} of Theorem~2 at hand, the proof is done via a suitable local comparison function.
Let $x_0\in \Omega$ be such that $f(x_0)=0$ and take $b_0\in(0,1)$ small such that
\be\label{inclx0}
\overline B(x_0,2b_0)\subset \Omega.
\ee
We consider the following function
\beaa
w(x,t):=1-A\bigl[\phi(x)+(T-t)\bigr]^{1/(p+1)}
\qh{ in $\overline B(x_0,b)\times[0,T)$},
\eeaa
where
$$ \phi(x):=\kappa b^2 \left(1-\frac{|x-x_0|^2}{b^2}\right)^2.$$
Here, $\kappa\in (0,1)$ and $b\in (0,b_0)$ are constants to be chosen later and $A$ is a fixed positive constant
such that $A\le\gamma b_0$ and $A\le (1+T)^{-1/(p+1)}$ (where $\gamma$ is the constant given in~\eqref{rate}).
Note that
$$w(x,0)=1-A[\phi(x)+T]^{1/(p+1)}\ge 0 \qh{ for $x\in \overline B(x_0,b)$}$$
and
\beaa
w(x,t)&=&1-A(T-t)^{1/(p+1)}\\
&\ge& 1-\gamma\delta(x)(T-t)^{1/(p+1)}\ge u(x,t) \qh{ for $(x,t)\in\partial B(x_0,b)\times(0,T)$,}
\eeaa
due to (\ref{inclx0}), $A\le\gamma b_0$ and \eqref{rate}.
We compute, in $B(x_0,b )\times(0,T)$,
\beaa
&&w_t- \Delta w-f(x)(1-w)^{-p}\\
&=&\frac{A}{p+1}\bigl[\phi(x)+(T-t)\bigr]^{-1+\frac{1}{p+1}} +\frac{A}{p+1}\bigl[\phi(x)+(T-t)\bigr]^{-1+\frac{1}{p+1}} \Delta\phi \\
&&\quad -\frac{Ap}{(p+1)^2}\bigl[\phi(x)+(T-t)\bigr]^{-2+\frac{1}{p+1}} |\nabla \phi|^2-f(x)A^{-p}\bigl[\phi(x)+(T-t)\bigr]^{-p/(p+1)}\\
&=&\frac{A}{p+1}\bigl[\phi(x)+(T-t)\bigr]^{-p/(p+1)}\\
&&\quad \times \left\{1+\Delta\phi-\frac{p}{p+1}\bigl[\phi(x)+(T-t)\bigr]^{-1} |\nabla \phi|^2-(p+1)A^{-p-1} f(x) \right\},\\
\eeaa
hence
\bea
&&w_t- \Delta w-f(x)(1-w)^{-p}\label{Parabw}\\
&\ge&\frac{A}{p+1}\bigl[\phi(x)+(T-t)\bigr]^{-p/(p+1)}
\left\{1 +\Delta\phi - \frac{p}{p+1}\frac{|\nabla \phi|^2}{\phi}-(p+1)A^{-p-1} f(x)\right\}. \notag
\eea
Moreover, in $B(x_0,b)$ we have
\beaa
&& \nabla\phi(x)=-4\kappa \left(1-\frac{|x-x_0|^2}{b^2}\right)\,(x-x_0),\\
&& \Delta\phi(x)=-4{n}\kappa \left(1-\frac{|x-x_0|^2}{b^2}\right)+8\kappa \left(\frac{|x-x_0|}{b}\right)^2
\ge -4{n}\kappa \\
&&\frac{|\nabla\phi(x)|^2}{\phi (x)} =16\kappa \left(\frac{|x-x_0|}{b}\right)^2 \le 16\kappa. \\
\eeaa
Now, since $f(x_0)=0$, we may choose $b>0$ small enough so that
$$(p+1)A^{-p-1}\sup_{x\in B(x_0,b)}f(x)\le \frac{1}{3}.$$
Then we can choose $\kappa {=\kappa(n)}\in(0,1)$ small enough such that
\beaa
\Delta\phi(x)\ge -\frac{1}{3}, \quad \frac{|\nabla\phi(x)|^2}{\phi (x)}\le\frac{1}{3}\quad\mbox{ for all $x\in B(x_0,b)$.}
\eeaa
Therefore, $w_t- \Delta w-f(x)(1-w)^{-p}\ge 0$ in $B(x_0,b )\times(0,T)$,
and it follows from the comparison principle that $u\le w$ in $B(x_0,b)\times(0,T)$.
{Since $\min_{\overline B(x_0,b/2)}\phi>0$,} this implies that $x=x_0$ is not a quenching point and the theorem is proved.
\qed
\bigskip
\section{Proof of Theorem 3}
(i) {\bf Case of (\ref{HypBdry1}).} Assume without loss of generality that $0\neq x_0\in\partial\Omega$ with $B(0,|x_0|)\cap\Omega=\emptyset$.
This (exterior ball condition) is possible due to the assumption that $\partial\Omega\in C^{2+\nu}$.
We look for a supersolution of the form $z(x)=1-C(d-r)^{\color{black} \beta}$, $r=|x|$, with $\beta>1$, $C{>0}, d>|x_0|$
to be chosen.
For $0<r<d$, we have:
\beaa
z_r&=&\beta C(d-r)^{\beta-1} \\
z_{rr}&=&-\beta (\beta-1)C (d-r)^{\beta-2}.
\eeaa
Set $\omega:=\Omega\cap\{x\in\BR^n;\ |x_0|<|x|<d\}$.
Choosing $\beta=2/(p+1)>1$, so that $\beta-2=-\beta p$, we compute in $\omega$:
\beaa
-\Delta z-f(x)(1-z)^{-p}&=&\beta (\beta-1)C(d-|x|)^{\beta-2}-\frac{\beta C(n-1)(d-|x|)^{\beta-1}}{|x|}\\
&&\ -f(x)\bigl[C(d-|x|)^\beta\bigr]^{-p} \\
&=&C(d-|x|)^{-\beta p}\Big[\beta(\beta-1)-\frac{\beta (n-1)(d-|x|)}{|x|}-C^{-p-1}f(x)\Bigr].
\eeaa
Next taking $d=d(p,n,|x_0|)>|x_0|$ close to $|x_0|$ and $C=C(p,\|f\|_\infty)>0$ large, we then have,
in $\omega$:
$$-\Delta z-f(x)(1-z)^{-p}\ge
C(d-|x|)^{-\beta p}\Big[\beta(\beta-1)-\frac{\beta (n-1)(d-|x_0|)}{|x_0|}-C^{-p-1}\|f\|_\infty\Bigr]\ge 0.$$
By taking $d$ possibly closer to $|x_0|$, we have
$$z(x)\ge 1-C(d-|x_0|)^\beta\ge 0\qh{ in $\overline\omega$},$$
hence also $z(x)\ge u(x,t)=0$ for $x\in \partial\omega\cap\{|x|<d\}=\partial\omega\cap\partial\Omega$ and $t\in (0,T)$.
Since on the other hand $z(x)=1>u(x,t)$ for $x\in \partial\omega\cap\{|x|=d\}$ and $0<t<T$, we deduce from the comparison principle that
$z\ge u$ on $\omega\times (0,T)$. Therefore $x_0$ is not a quenching point.
\medskip
(ii) {\bf Case of (\ref{HypBdry}).} The proof {relies on estimate \eqref{rate} and} is a modification of that of Theorem~1.
Let $\Omega_\eta=\{x\in\Omega;\ \delta(x)<\eta\}$. {\color{black} There exists $\eta_0>0$ such that
$\Omega_\eta$ is a smooth bounded domain for all $\eta\in (0,\eta_0)$, due to $\Omega$ being a smooth domain.} We have
$\partial\Omega_\eta=\partial\Omega\cup\Gamma_\eta$, where $\Gamma_\eta=\{x\in\Omega;\ \delta(x)=\eta\}$.
Let $\psi_\eta$ be the unique solution of the problem
\beaa
&&\Delta \psi_\eta=0,\quad x\in\Omega_\eta, \\
&&\psi_\eta=1, \quad x\in \partial\Omega, \\
&&\psi_\eta=0,\quad x\in \Gamma_\eta. %
\eeaa
The function $\psi_\eta$ is smooth and, by the strong maximum principle,
we have $0<\psi_\eta<1$ in $\Omega_\eta$.
Letting
$$\phi=k\psi_\eta^2,$$
we consider the function
\beaa
w(x,t):=1-\gamma\eta\bigl[\phi(x)+(T-t)\bigr]^{1/(p+1)}
\qh{ in $\overline\Omega_\eta\times [0,T)$},
\eeaa
where $k\in (0,1)$ and $\eta\in (0,\eta_0)$ are constants to be chosen later {\color{black} and $\gamma$ is the constant given in \eqref{rate}.}
First assuming $\eta\le \eta_1:=\min\bigl(\eta_0,\gamma^{-1}(1+T)^{-1/(p+1)}\bigr)$, we have
$$w(x,0)=1-\gamma\eta[\phi(x)+T]^{1/(p+1)}\ge 0 \qh{ in $\overline\Omega_\eta,$}$$
along with $w(x,t)\ge 0$ on $\partial\Omega\times (0,T)$ and, by \eqref{rate},
\beaa
w(x,t)&=&1-\gamma\eta(T-t)^{1/(p+1)}= 1-\gamma\delta(x)(T-t)^{1/(p+1)}\ge u(x,t) \qh{ on $\Gamma_\eta\times(0,T)$.}
\eeaa
Formula (\ref{Parabw}) remains valid, with $A$ replaced by $\gamma\eta$,
and moreover we have
$$\Delta\phi=2k|\nabla\psi_\eta|^2 \qh{ and }\quad \frac{|\nabla\phi|^2}{\phi}=4k|\nabla\psi_\eta|^2.$$
Therefore,
\beaa
&&w_t-\Delta w-f(x)(1-w)^{-p}\\
&\ge&\frac{\gamma\eta}{p+1}\bigl[\phi(x)+(T-t)\bigr]^{-p/(p+1)}
\left\{1-\frac{4kp}{p+1} |\nabla \psi_\eta|^2-(p+1)(\gamma\eta)^{-p-1} f(x)\right\}
\eeaa
{in $\Omega_\eta\times [0,T)$.} Now, by assumption (\ref{HypBdry}), we may choose $\eta\in(0,\eta_1)$ small enough so that
$$(p+1)(\gamma\eta)^{-p-1}\sup_{x\in\Omega_\eta}f(x)\le \frac{1}{2}.$$
Then we can choose $k=k(\eta)>0$ small enough so that
\beaa
4k|\nabla\psi_\eta|^2\le\frac{1}{2}\quad\mbox{ for all $x\in\Omega_\eta$.}
\eeaa
Therefore, $w_t-\Delta w-f(x)(1-w)^{-p}\ge 0$ in $\Omega_\eta\times(0,T)$,
and it follows from the comparison principle that $u\le w$ in $\Omega_\eta\times(0,T)$.
Since $\min_{\overline\Omega_{\eta/2}}\psi_\eta>0$, this guarantees that no quenching occurs near the boundary.
\qed
\medskip
{{\bf Remark.} Although it is not clear if Theorem 3(i) has a direct relation with this fact,
it is interesting to recall that, when $f$ is constant, quenching for problem (\ref{pdeGen})-(\ref{icGen}) is
incomplete (in the sense of existence of a suitable weak continuation after $t=T$)
if and only if $0<p<1$ (cf.~\cite{Phi87, FLV, GV97}).}
|
\section{Introduction}
In the active Sun, the emergence of magnetic flux from the solar interior can lead to the formation of ARs
\citep[e.g.][]{zwaan85}, which are often associated with coronal mass ejections (CMEs), flaring activity and
other dynamic events \citep[see e.g.][]{vandriel95,schrijver09,arcr12}.
In adittion, observational studies \citep[see e.g.][]{hudson98,canfield99} have revealed that
ARs with S-shaped morphology are the most favourables candidates for eruptive activity in the Sun.
The term ``sigmoid'' has been used by \cite{rust96} to denote the overal (forward or reverse)
S-like shape structure of an AR.
Observations \citep[see e.g.][]{sterling00,pevtsov02,liu10} have also shown that eruptive ARs, which initially
display a degree of twist adopting a sigmoidal
shape, evolve into a post-eruption flare arcade that consists of fieldlines with a cusp-like shape. This process is
known as ``sigmoid-to-arcade'' evolution and it could happen repeatedly \citep[e.g.][]{gibson02,nitta01} in some ARs: a bright (EUV/X-ray) sigmoid is reformed
after each CME(-like) eruption, which is accompanied by a hot arcade underneath it.
Numerical experiments have demonstrated the importance of magnetic flux emergence on driving CME-like eruptions associated with
flaring activity in emerging flux regions (EFRs) \citep[e.g.][]{shibata_r11, kusano12}. Simulations have also shown the formation of sigmoidal structures
in EFRs in conjuction
with the (partial) eruption of magnetic flux ropes \citep[e.g.][]{gibson06,arc09}. The onset of recurrent eruptions has been studied in the context
of a break-out magnetic scenario \citep[e.g.][]{devore08} and in a smimilar manner in EFRs
(e.g. reconnection between the emerging and a pre-existing magnetic field, \citep[e.g.][]{arc08,mactaggart09}).
The repeated ``sigmoid-to-arcade'' evolution associated with CMEs has been shown by the kinematically driven (quasi-static) emergence of a highly
twisted magnetic torus into a preexisting potential coronal field \citep[]{chat13}. In this Letter, we present a 3D MHD experiment
of the onset of recurrent CME-like eruptions driven by the {\it dynamical} emergence of a horizontal twisted flux tube in a highly
stratified atmosphere. We find that the ``sigmoid-to-arcade'' process occurs naturally due to reconnection of sheared fieldlines during the eruptions.
\section{The model}
We solve the MHD equations in
Cartesian geometry, using the Lare3d code \citep[]{arber01}. Viscous and
Ohmic heating are considered through shock viscosity and Joule dissipation.
Uniform explicit resistivity is included, with $\eta=10^{-3}$.
Initially, the plasma is embedded into a plane-parallel hydrostatic atmosphere.
A sub-photospheric adiabatically stratified layer resides in the range ($-7.2\,\mathrm{Mm} \leq z < 0\,\mathrm{Mm}$).
The layer above, at $0\,\mathrm{Mm}\leq z < 2.3\,\mathrm{Mm}$, which
is isothermal ($5100\,\mathrm{K}$) at the beginning and then the temperature increases with height up to $\approx 4\times10^{4}\,\mathrm{K}$,
is mimicking the photosphere/chromosphere.
The layer at $2.3\,\mathrm{Mm}\leq z \leq 3.1\,\mathrm{Mm}$ represents the transition region.
An isothermal coronal layer ($\approx 1\,\mathrm{MK}$) is included at $3.1\,\mathrm{Mm} < z \leq 57.6\,\mathrm{Mm}$.
The initial magnetic field is a horizontal twisted flux tube at $z_{0}=-2.1\,\mathrm{Mm}$ and
oriented along the positive $Y$-axis. The axial field of the cylindrical tube is defined by
\begin{equation}
B_y=B_0\,\mathrm{exp}(-r^2/R^2), \,\,\,\,\, B_\phi=\alpha\,r\,B_y,
\end{equation}
where $R=450\,\mathrm{km}$ is the radius of the tube, $r$ is the radial distance
from the tube axis ($r^{2} = x^{2} + (z + z_{0})^{2}$) and $\alpha=0.0023\mathrm{km^{-1}}$ is the uniform twist
around the axis of the tube, which is stable to the kink instabilty. A density deficit is applied along the axis of the tube, making
its central part more buoyant than its footpoints:
\begin{equation}
\Delta\,\rho=[p_t(r)/p(z)]\,\rho(z)\,\mathrm{exp}\,(-y^2/\lambda^2),
\end{equation}
where $p_t$ is the pressure within the flux tube and $\lambda$ defines the length of the buoyant part of the tube.
We use $\lambda=0.9\,\mathrm{Mm}$ and an initial field strength for the tube that corresponds to plasma ${\beta}\approx14$.
The numerical domain
is $[-32.4,32.4] \times [-32.4,32.4] \times[-7.2,57.6]\,\mathrm{Mm}$ in the longitudinal ($x$), transverse
($y$) and vertical ($z$) directions, respectively. The grid has 420 nodes in all
directions with periodic boundary conditions in $y$. Open boundary conditions have been implemented along $x$
and at the top of the numerical domain, allowing outflow of plasma.
The bottom boundary is a non-penetrating, perfectly conducting wall.
\section{Results and discussion}
\begin{figure}[t]
\centering
\includegraphics[width=1.\linewidth]{./f1.pdf}\caption{Temperature and magnetic field topology during the eruptions.
Shown are the two vertical midplanes and the horizontal slice at the base of the photosphere.}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[width=0.9\linewidth]{./f2.pdf}\caption{ Field line topology during the first eruption (panels (a)-(c)).
The horizontal slice shows the Bz distribution at the photosphere (red;positive, blue;negative within the range $[-500,500]\,\mathrm{G}$.
Times are $t=62.8\,\mathrm{min}$,
$t=68.5\,\mathrm{min}$ and $t=71.4\,\mathrm{min}$ for panels (a)-(c) respectively. Logarithmic density and temperature, and vertical velocity distribution at the
vertical (xz) midplane are shown in panels (d)-(f) at $t=71.4\,\mathrm{min}$.}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[width=0.9\linewidth]{./f3.pdf}\caption{{\bf Top:} Temporal evolution of the magnetic (black line ) and
kinetic (red) energy above the photosphere. {\bf Bottom:} Temporal evolution of the maximum positive/negative (red/blue resepectively) $V_\mathrm{z}$ and
maximum temperature (black) above the photosphere.}
\end{figure}
\begin{figure*}[t]
\centering
\includegraphics[width=0.9\linewidth]{./f4.pdf}\caption{The sigmoid-to-arcade transformation during the first CME-like eruption
(panels a-c) and the reappearance of a sigmoid before the second eruption (panel d).
The value of the isosurface represents $\ge 50\%$ (panels a,b,d) and $\ge 25\%$
(panel c) of the maximum value of ${J\cdot B}/{\bf B}$ above the transition region. The horizontal slice shows the Bz distribution at the photosphere
(red/yellow;positive, blue/pink;negative within the range $[-450,450]\,\mathrm{G}$.)
Times are $t=65.7\,\mathrm{min}$, $t=68.5\,\mathrm{min}$, $t=71.4\,\mathrm{min}$ and $t=77.2\,\mathrm{min}$ for panels (a)-(d) respectively.}
\end{figure*}
Our experiment shows the recurrent eruptive behaviour of an EFR during 4.5 hours of its evolution.
Figure 1 shows the temperature distribution together with magnetic fieldlines, which have been traced from the
footpoints of the sub-photospheric flux tube. Panel 1(a) shows the cool adiabatic expansion
of the emerging field. The outermost fieldlines have already emerged into the corona forming an envelope (ambient) field for the
magnetic flux, which continues to emerge from the solar interior. Part of this magnetic flux forms a new magnetic flux rope that erupts into the
outer solar atmosphere in an ejective manner opening the envelope field (Panels 1(b),1(c)). There are at least four ejective
eruptions, during the evolution (three of them are shown in Figure 1) of the EFR, displaying several common characteristics such as: a) the marked expansion
of the erupting magnetized volume, b) the transport of dense (low-atmospheric) material to the high atmosphere, c) the explosive heating of the plasma underneath
the erupting core of the field and d) the formation of a hot arcade with a cusp-like shape in the low atmosphere.
The formation of a new flux rope and the driving mechanism of its eruption in similar flux emergence experiments have been studied in
previous simulations \citep[e.g.][]{maglon01, man04, arc08b, arc12}. Figure 2 (panels a-c) is a synoptic illustration of the formation
and rise of the first new flux rope in the present experiment. At $t=62.8\,\mathrm{min}$ (2(a)), the emerging fieldlines (blue and cyan), which have
previously undergone shearing,
reconnect with each other to form two new sets of fieldlines: the white fieldlines that (will) constitute the central part of the developing flux rope and
the yellow fieldlines that stay behind and do not erupt. The red fieldlines represent
the envelope field. At $t=68.5\,\mathrm{min}$ (2(b)), the white fieldlines are traced
from the center of the cross section of the erupting flux rope, at the vertical (xz) midplane. The blue and cyan fieldlines continue to reconnect,
forming new fieldlines: the orange fieldlines, which are wrapping around the footpoints of the white fieldlines developing a
magnetic flux rope structure and the yellow fieldlines, which form an arcade. Notice that now the envelope fieldlines are stretched out
so that they are about to reconnect in a tether-cutting manner underneath the flux rope. At $t=71.4\,\mathrm{min}$ (2(c)), the well-developed flux rope
consists of straight (white) fieldlines along its axis and twisted fiedlines (blue/cyan/orange/red) around its axis.
An important result is that, at this stage of the evolution, the enevelope fieldlines reconnect in a tether-cutting manner at a strong and thin current layer
under the rope.
This process triggers the ultimate ejective eruption of the flux rope towards the outer solar atmosphere. Therefore, in this case, the release of
the downward tension of the
envelope fieldlines, which under certain conditions can halt the ejective eruption \citep[e.g.][]{arc12}, is not released by external reconnection
with a pre-existing
coronal magnetic field \citep[e.g.][]{arc08b} in a break-out manner but by tether-cutting internal
reconnection underneath the erupting plasma.
Panel 2(d) shows that the eruption of the flux rope brings dense plasma from the low atmosphere into the corona. The heavy material ($\approx 1-2$ orders heavier
than the background plasma) is accumulated
at the dips of the twisted fieldlines of the flux rope. Panel 2(e) shows that the ejective phase of the eruption is characterized by the cool adiabatic
expansion of the field and the heating of the plasma due to tether-cutting reconnection of the fieldlines.
Panel 2(f) shows that high-speed ($\approx 800\,\mathrm{km\,s^{-1}}$) bi-directional jets are emitted vertically (upward/downwad) from the
reconnection site underneath the rising flux rope.
The upward reconnection jet adds momentum to the rising flux rope and heats the
plasma inside the erupting volume. Therefore (panel 2e), cool (purple) and hot (yellow/red) plasma is found to be expelled together with the erupting
core of the flux rope. In the first two eruptions (which are faster, see also Figure 3), the top edge of the upward jet nearly reaches
to the core of the erupting field. Due to its high speed, it deforms the concave part of the twisted fieldlines that
surrounds the erupting core from below. The downward reconnection jet collides with the top (flux pile-up) regime of the arcade
forming a termination shock. As a result, the plasma is compressed locally at high temperatures, causing flaring of the arcade.
Eventually, a second flux rope is formed at the upper photosphere/chromosphere, as new fieldlines emerge and undergo shearing and reconnection
(such as the blue and cyan fieldlines in panels 2(a)-2(c)).
The innermost fieldlines of the arcade udergo the same process and, thus, they
are incorporated in the formation of the flux rope. The outermost fieldlines of the arcade constitute a new envelope field for the second
flux rope to rise into. Initially, the flux rope
rises slowly (with a speed of about $40\,\mathrm{km\,s^{-1}}$ until $t=80\,\mathrm{min}$) and then it erupts in a fast-rise phase, accelerating to a velocity of $\approx 280\,\mathrm{km\,s^{-1}}$
in the next $90\,\mathrm{sec}$.
The onset of acceleration is accompanied by tether-cutting reconnection of the new envelope's fieldlines.
The following eruptions occur in a slightly different manner.
The magnetic field at the {\it center} of the EFR
does not become strong enough to emerge above the photosphere with the standard $\Omega$-loop like configuration.
On the contrary, it spreads out horizontally and the magnetic pressure increases in the
vicinity of the two main polarities of the EFR. There, the magnetic fieldlines can emerge forming two magnetic lobes, similar to the emergence of
a weakly twisted emerging field \citep[]{arc13}. The lateral expansion of the magnetic lobes bring their fieldlines into contact at the center of the EFR and
reconnect, forming another flux rope. The eruption of the flux rope is driven by the Lorentz force, which is eventually enhanced by tether-cutting
reconnection of the envelope fieldlines.
In summary, the tether-cutting reconnection following the first eruption restructures the
magnetic configuration in the solar atmosphere, so that the post-eruption state resembles the
pre-eruption state of the first eruptive event (Figure 2(a)). Eventually, emergence of new magnetic
flux and associated shearing, inject free magnetic energy into the system initiating a second
eruption very similar (homologous) to the first one.
Overall, the sequence of the homologous eruptions is attributed to a) ongoing flux emergence,
shearing and reconnection in the low atmosphere (e.g. up to the transition region) and b) to
the tether-cutting reconnection in the corona and the dynamical reconfiguration of the system
to a state similar to the initial one.
There are two differences in the topology of the field and the dynamics prior to the eruptions. Firstly, the
envelope field in the initial eruption consists of the uppermost emerging fieldlines, while in
the following eruptions it is formed by the fieldlines of the flare arcade. Secondly, the
formation of the erupting flux rope in the first two eruptions occurs due to shearing and reconnection
of the emerging field at the center of the AR. In the other two eruptions, the emerging field
develops two side magnetic lobes (see the work by \citet[]{arc13}), which eventually reconnect, forming
the magnetic flux rope.
Despite these differences, the key processes which are responsible for the driving
of the homologous eruptions in our experiment are similar in all events.
Figure 3 (top) shows the temporal evolution of the total magnetic and kinetic energy above the photosphere. The four eruptions start to occur at the
times marked by the vertical dashed lines. In each eruption, magnetic energy drops and kinetic energy increases rapidly. The first increase of the
kinetic energy ($t\approx 40\,\mathrm{min}$) corresponds to the initial emergence of the magnetic field. The second eruption starts while the lower part (i.e. where
tether cutting reconnection of the envelope fieldlines occurs) of the first eruption is still within the numerical domain.
For example, at $t\approx 83\,\mathrm{min}$
the tether cutting reconnection upflow (jet) of the first eruption originates at $z\approx 44\,\mathrm{Mm}$ and, thus, the
corresponding total kinetic energy resembles the
accumulated energy of the first two eruptions. After the first eruption, the decrease (increase) of the magnetic (kinetic) energy becomes lower with
each of the following eruptions, which indicates that the CME-like eruptions become progressively less energetic. This is because there is a certain
amount of flux and energy available to the system for driving the eruptions. This energy comes from the sub-photospheric magnetic flux tube.
Due to dynamical emergence, the flux is eventually exchausted and, thus, the eruptions produced are less energetic.
Notice that the total magnetic energy increases after each eruption. This increase is important for the build-up of the neccessary amount of energy that is
required for the onset of the successive eruptions. In the first and second eruption, the increase is due to the dynamical emergence and
the shearing of the magnetic field around the center of the EFR. In the following eruptions, shearing is less pronounced but emergence continues to occur, which
leads to the formation of the two magnetic lobes. Eventually, the system reaches a quasi-static equilibrium and the magnetic energy saturates.
Figure 3 (bottom) shows the temporal evolution of the vertical component of the velocity field (maximum and minimum value) and the temperature (maximum) above
photosphere. Strong bi-directional flows are emitted during each eruption. These flows are the reconnection jets, which originate in the
current sheet underneath the erupting flux rope. There is a very good correlation between the upflows and the downflows during the evolution of the
system: for every upflow there is a corresponding downflow with the same almost magnitude. The fastest upflow occurs in the first eruption ($\approx
800\,\mathrm{km\,s^{-1}}$). In the other eruptions, the associated jets run with progressively lower speeds, within the range $450-350\,\mathrm{km\,s^{-1}}$.
The measured velocity
at the center of the erupting flux ropes, in the last snapshot before they exit the domain and while they are accelerated, is
between $\approx 200-250\,\mathrm{km\,s^{-1}}$. This indicates that indeed the reconnection jets can add momentum to the erupting plasma volume.
The evolution of the temperature shows that each eruptive event is followed by intense plasma heating. In a similar manner to the temporal evolution
of $V_\mathrm{z}$, the maximum heating decreases over time. It drops from $17\,\mathrm{MK}$ after the first eruption to about $10\,\mathrm{MK}$ in the
last eruptive event. Therefore,
the heating decreases when the eruption is less energetic. The intense heating occurs at the apex of the cusp-like arcade underneath the
reconnection site of the produced jets.
Therefore, each eruption is followed by a flaring episode. The time offset between the onset of the eruption
and the heating of the arcade is determined by the
time that the downward reconnection jet takes to reach the arcade and heat it.
The recurrent CME-like eruptions are associated with the appearance of sigmoidal structures in the EFR. Figure 4 shows the
three-dimensional visualization of the isosurface ${J\cdot B}/{\bf B}$, where J is the electric current and B is the magnetic field strength.
At $t=65.7\,\mathrm{min}$ (panel (a)), the isosurface and the fieldlines that are traced from this structure adopt an overal S-like shape associated with the twist and writhe
of the rising magnetic field. Some of these fieldlines (white) are dipped at the center of the sigmoid and expand towards its two ends (``elbows'').
There are also sheared fieldlines (blue/cyan) that surround only one of the ``elbows'' and, thus, they make only a half turn along the sigmoid.
As these fieldlines come closer together, the current becomes large at their interface (along the PIL) developing the sigmoidal structure.
Eventually, the blue and cyan fieldlines will reconnect around the center of the EFR forming a new flux rope. Due to reconnection, we expect that the
sigmoid's fieldlines will be heated to high temperatures. At this stage of the simulation, we find that the dense plasma of the sigmoidal structure has an
average temperature of the $O(5\,\times 10^{5}) {K}$ over the integrated volume (i.e. above the transition region) of the isosurface and, thus, it may
appear as a bright S-shaped structure at the solar atmosphere.
At $t=68.5$ (panel (b)), the sigmoid develops a loop-like structure at its center. This is the current-currying flux rope that starts to rise, leading to the
partial eruption of the sigmoid. The (white) fieldlines, which have been traced from the loop, belong to the erupting flux rope. The red fieldlines have
been traced from the center of the sigmoidal isosurface underneath the flux rope.
The flaring of the plasma, which follows the eruption, starts from this central area of the sigmoid.
As the eruption proceeds (panel (c)), a new thin and curved current layer is formed underneath the erupting plasma. This
is where tether-cutting reconnection occurs (TC current).
Thus, fieldlines (orange), which were previously part of the enevelope field, now reconnect adopting a V-shape at
the center of the current layer.
Under the TC current, the downward reconnected fieldlines (red)
form the ``flare'' arcade, which is heated to more than $15\,\mathrm{MK}$. Thus, the bright arcade does not appear to be part of the
pre-flare sigmoid but it is a new structure associated with the tether-cutting reconnection. The (blue/cyan) fieldlines are traced from the lower-lying
remnants of the sigmoid, which temporarily fades away. However, shortly after the first eruption (panel (d)), the dynamical emergence
and the shearing of the field continue to operate, leading to reappearance of a sigmoidal structure (blue/cyan/white fieldlines have been traced
in a similar manner to panel a).
A similar process (i.e. sigmoid-to-arcade conversion) is repeated during the following eruptions.
We have performed a preliminary parametric study, which shows that the parameters of the system may affect its evolution and dynamics.
For instanse, for larger values of plasma $\beta$ or $\lambda$ of the sub-photospheric flux tube the eruptions occur at later times. In adittion, we find that
for an initial field strength of the tube that correrponds to $\beta \geq 39$, the eruptions are confined.
The ejection of helicity and the reformation of sigmoids for a wider range of parameters
will be presented in forthcoming work.
\acknowledgments
The simulations were performed on the STFC and SRIF funded
UKMHD cluster, at the University of St Andrews. The authors acknowledge support by EU (IEF-272549 grant) and
the Royal Society.
V.A and A.W.H are grateful
for in-depth discussions during the ISSI workshop: ``Magnetic flux emergence in the solar atmosphere'' in Bern.
\bibliographystyle{apj}
|
\section{Noncrossing arc diagrams}\label{diagrams sec}
The key objects in this paper are \newword{noncrossing arc diagrams}, certain diagrams consisting of arcs satisfying certain rules, including the requirement that arcs not cross.
Each diagram begins with $n$ distinct points on a vertical line.
We identify the points with the numbers $1,\ldots,n$, with $1$ at the bottom.
Each diagram is consists of some (or no) curves called \newword{arcs} connecting the points.
Each arc must satisfy the following requirement:
\begin{enumerate}
\item[(A)]
The arc connects a point $p$ to a strictly higher point $q$, moving monotone upwards from $p$ to $q$ and passing either to the left or to the right of each point between $p$ and $q$.
The arc may pass to the left of some points and to the right of others.
\end{enumerate}
The diagram must also satisfy the following two pairwise compatibility conditions:
\begin{enumerate}
\item[(C1)]
No two arcs intersect, except possibly at their endpoints.
\item[(C2)]\setcounter{enumi}{2} \label{diag def}
No two arcs share the same upper endpoint or the same lower endpoint.
\end{enumerate}
\begin{figure}[ht]
\scalebox{0.9}{\includegraphics{diagrams3}}
\caption{Noncrossing arc diagrams on $3$ points}
\label{diagrams3}
\end{figure}
Each noncrossing arc diagram determines some combinatorial data, namely which pairs of points are joined by an arc and which points are left and right of each arc.
Two noncrossing arc diagrams are considered to be combinatorially equivalent if they determine the same combinatorial data, and we consider arcs and noncrossing arc diagrams only up to combinatorial equivalence.
For $n=1$, there is one noncrossing arc diagram (with no arcs), and for $n=2$ there are two noncrossing arc diagrams (with one arc or no arcs).
Figures~\ref{diagrams3} and~\ref{diagrams4} show the $6$ noncrossing arc diagrams on $3$ points and the $24$ noncrossing arc diagrams on $4$ points.
\begin{figure}
\scalebox{0.82}{\includegraphics{diagrams4}}
\caption{Noncrossing arc diagrams on $4$ points}
\label{diagrams4}
\end{figure}
Say two arcs are \newword{compatible} if there exists a noncrossing arc diagram containing those two arcs.
Compatibility is a combinatorial condition, depending only on the endpoints of the two curves and on which points are left and right of each curve:
Given arcs $\alpha_1$ and $\alpha_2$, suppose there is a point $p$ that is left of (or an endpoint of) $\alpha_1$ and is right of (or an endpoint of) $\alpha_2$, with $p$ not an endpoint of both arcs.
Then a noncrossing arc diagram containing $\alpha_1$ and $\alpha_2$ must have $\alpha_1$ to the right of $\alpha_2$.
The two curves are compatible if and only they don't share upper endpoints or lower endpoints and if there do not exist both a point $p$ forcing $\alpha_1$ to be right of $\alpha_2$ and a point $p'$ forcing $\alpha_2$ to be right of $\alpha_1$.
Given any noncrossing arc diagram, the arcs in the diagram are pairwise compatible.
However, \textit{a priori} we don't know, given a collection of pairwise compatible arcs, if it is possible to fix a representative for each arc so that the \emph{representatives} satisfy (C1) pairwise.
In Section~\ref{proof sec}, we prove that it \emph{is} possible to fix such a set of representatives.
One family of noncrossing arc diagrams is very familiar.
A \newword{left arc} is an arc that does not pass to the right of any point, and we will call a noncrossing arc diagram having only left arcs a \newword{left noncrossing arc diagram}.
The left noncrossing arc diagrams are more commonly known as \newword{noncrossing partitions}, although it is more typical to draw points on a vertical line and allow only arcs that don't pass below any points.
The enumeration of noncrossing partitions is well-known, and we give it here in the language of this paper:
The number of left noncrossing arc diagrams on $n$ points is the Catalan number $C_n=\frac1{n+1}\binom{2n}n$.
The number of such diagrams with $k$ arcs is the Narayana number $\frac1n\binom n k\binom n{k-1}$.
One might similarly define a \newword{right arc} to be an arc that does not pass to the left of any point, and the same enumerative statements hold with ``left'' replaced by ``right.''
Surprisingly, something nice also happens when we mix right arcs with left arcs:
\begin{theorem}\label{diagrams left right}
The number of noncrossing arc diagrams on $n$ points having only left arcs and right arcs is the Baxter number
\[B(n)=\binom{n+1}{1}^{-1}\binom{n+1}{2}^{-1}\,\,\sum_{k=0}^{n-1}\binom{n+1}{k}\binom{n+1}{k+1}\binom{n+1}{k+2}.\]
\end{theorem}
Among several combinatorial objects already known to be counted by the Baxter number $B(n)$ are the \newword{diagonal rectangulations} with $n$ rectangles.
(This is due to \cite{ABP,CGHK}. See \cite[Section~1]{rectangle} and \cite[Remark~6.6]{rectangle} for details on attribution.)
Fixing a square and one diagonal of the square, diagonal rectangulations are (combinatorial types of) decompositions of the square into rectangles such that each rectangle's interior intersects the given diagonal.
A larger set of rectangulations are the \newword{generic rectangulations}.
(See~\cite{generic}.)
These are (combinatorial types of) decompositions of a square into rectangles such that no four rectangles have a common corner.
Just as the diagonal rectangulations are in bijection with a certain set of noncrossing arc diagrams by Theorem~\ref{diagrams left right}, the generic rectangulations are in bijection with a larger set of noncrossing arc diagrams.
An \newword{inflection} of an arc in a diagram is a pair of adjacent points with one left of the arc and the other right of the arc.
If we draw arcs in the natural way as in Figures~\ref{diagrams3} and~\ref{diagrams4}, the arc has an inflection point (in the sense of curvature) between the two points comprising the inflection.
The noncrossing arc diagrams of Theorem~\ref{diagrams left right} are diagrams whose arcs have no inflections.
A similarly nice thing happens if we allow up to one inflection in each arc.
\begin{theorem}\label{diagrams inflection}
The noncrossing arc diagrams on $n$ points with each arc having at most one inflection are in bijection with generic rectangulations with $n$ rectangles.
\end{theorem}
Noncrossing partitions, diagonal rectangulations, and generic rectangulations are all in bijection with certain pattern-avoiding permutations.
These connections and the evidence for small $n$ suggest the following theorem, which is somewhat surprising \emph{a priori}.
\begin{theorem}\label{diagrams perm}
There are $n!$ noncrossing arc diagrams on $n$ points.
\end{theorem}
We prove Theorems~\ref{diagrams left right}, \ref{diagrams inflection}, and~\ref{diagrams perm} and several other enumerative statements in Sections~\ref{proof sec} and~\ref{quot sec}.
Before doing that, we explain canonical join representations of permutations in Section~\ref{can sec}.
Canonical join representations are not really necessary for the proofs of Theorems~\ref{diagrams left right}, \ref{diagrams inflection}, and~\ref{diagrams perm}.
However, canonical join representations explain ``what's really going on'' in these theorems and in many of the other results.
Furthermore, noncrossing arc diagrams provide the answer to a very natural question about canonical join representations (Question~\ref{can q}).
Noncrossing arc diagrams are foreshadowed in the work of Bancroft~\cite{Bancroft} and Petersen~\cite{Petersen} on shard intersections in the Coxeter arrangement of type A.
\section{Canonical join representations of permutations}\label{can sec}
A \newword{join representation} for an element $x$ in a finite lattice $L$ is an identity $x=\Join S$, where $S$ is a subset of $L$.
The join representation $x=\Join S$ is \newword{irredundant} if there is no proper subset $S'\subset S$ such that $x=\Join S'$.
In particular, if $\Join S$ is irredundant, then $S$ is an antichain.
We define a relation $\ll$ on subsets of $L$ by setting $S\ll T$ if, for every $s\in S$, there exists a $t\in T$ with $s\le t$.
The relation $\ll$ restricts to a partial order on antichains (equivalent to containment order on order ideals generated by the antichains).
A join representation $x=\Join S$ is called the \newword{canonical join representation} of $x$ if it is irredundant and if every join representation $x=\Join T$ has $S\ll T$.
In other words, $S$ is the unique minimal antichain, in the partial order $\ll$, among antichains joining to $x$.
If $x=\Join S$ is the canonical join representation of $x$, then the elements of $S$ are called the \newword{canonical joinands} of $x$.
We sometimes abuse terminology by referring to the set $S$ itself, rather than the expression $x=\Join S$, as the canonical join representation of~$x$.
If an element has a canonical join representation, then each canonical joinand is join-irredicible.
That is, each canonical joinand $j$ covers exactly one element of $L$, or equivalently, there is no join representation for $j$ consisting of elements strictly below $j$.
The following proposition is immediate.
\begin{prop}\label{ji can}
An element $j$ of a finite lattice $L$ is join-irreducible if and only if its only canonicial joinand is $j$ itself.
\end{prop}
A lattice in which every element has a canonical join representation is called join-semidistributive. If the dual condition (every element has a canonical meet representation) holds as well, then the lattice is called \newword{semidistributive}.
The usual definition of semidistributivity involves two dual weakenings of the distributive laws, but the present definition is equivalent by \cite[Theorem~2.24]{FreeLattices}.
Suppose $L$ is a finite join-semidistributive lattice.
Not every antichain of join-irreducible elements in $L$ is a canonical join representation.
We define the \newword{canonical join complex} of $L$ to be the abstract simplicial complex whose vertices are the join-irreducible elements of $L$ and whose faces are the sets $S$ such that $\Join S$ is a canonical join representation.
This is a simplicial complex in light of Proposition~\ref{ji can} and the following proposition.
\begin{prop}
Suppose $L$ is a finite lattice and let $S$ be a subset of $L$.
If $\Join S$ is a canonical join representation and $S'\subseteq S$ then $\Join S'$ is also a canonical join representation.
\end{prop}
\begin{proof}
Let $x$ be the element $\Join S$, let $x'$ be the element $\Join S'$, and write $C$ for $S\setminus S'$.
Suppose $\Join T'$ is a join representation for $x'$ and let $s\in S'$.
We need to show that there exists $t\in T'$ with $s\le t$.
Writing $T=T'\cup C$, the expression $\Join T$ is a join representation for $x$.
Since $\Join S$ is the canonical join representation for $x$, there exists $t\in T$ such that $s\le t$.
Since $S$ is an antichain and $s\le t$, we see that $t\not\in C$.
Thus $t\in T'$ as desired.
\end{proof}
A permutation $x$ of $\set{1,\ldots,n}$ is a sequence $x_1x_2\cdots x_n$ such that $\set{x_1,\ldots,x_n}=\set{1,\ldots,n}$.
The weak order on permutations is a partial order whose cover relations are $x_1\cdots x_n\lessdot y_1\cdots y_n$ whenever there exists $i$ such that $x_i=y_{i+1}<y_i=x_{i+1}$ and such that $x_j=y_j$ for $j\not\in\set{i,i+1}$.
We write $S_n$ for the set of permutations of $\set{1,\ldots,n}$ partially ordered under the weak order.
Figure~\ref{weak ex} shows this partial order for $n=3$ and for $n=4$.
\begin{figure}
\scalebox{0.8}{\includegraphics{weakS3}}\qquad\qquad\scalebox{0.8}{\includegraphics{weakS4}}
\caption{The weak order on $S_n$ for $n=3$ and for $n=4$}
\label{weak ex}
\end{figure}
An \newword{inversion} in $x_1\cdots x_n$ is a pair $(x_i,x_j)$ with $i<j$ and $x_i>x_j$.
For example, the inversions of $25314$ are $(2,1)$, $(3,1)$, $(5,1)$, $(5,3)$ and $(5,4)$.
Write $\operatorname{inv}(x)$ for the \newword{inversion set} (the set of inversions) of $x$.
The weak order on permutations is characterized by containment of inversion sets.
That is, $x\le y$ in the weak order if and only if $\operatorname{inv}(x)\subseteq\operatorname{inv}(y)$.
The weak order on permutations is a lattice, and Duquenne and Cherfouh showed that it is semidistributive \cite[Theorem~3]{DuCh}.
(More generally, the weak order on any finite Coxeter group is semidistributive \cite[Lemme~9]{Poly-Barbut}.)
In Theorem~\ref{perm can join}, below, we describe canonical join representations of permutations explicitly, in particular giving an independent proof that the weak order on permutations is semidistributive.
The argument in \cite{DuCh} is very different, but a proof that is similar in spirit to the proof here can be obtained by combining \cite[Proposition~6.4]{con_app} with special cases of \cite[Lemma~3.5]{shardint} and \cite[Theorem~3.6]{shardint},
A \newword{descent} in a permutation is a pair $x_i$, $x_{i+1}$ of adjacent entries such that $x_i>x_{i+1}$.
Given a permutation $x_1\cdots x_n$ and a descent $x_i>x_{i+1}$, define $\lambda(x,i)$ to be the permutation given by $1,\ldots,(x_{i+1}-1)$, then the values $\set{x_j:j<i,\,x_{i+1}<x_j<x_i}$ in increasing order, then $x_i$, then $x_{i+1}$, then $\set{x_j:i+1<j,\,x_{i+1}<x_j<x_i}$ in increasing order, and finally the values $(x_i+1),\ldots,n$.
Thus $\lambda(x,i)$ is join-irreducible and covers the permutation obtained by swapping the adjacent entries $x_i$ and $x_{i+1}$ in $\lambda(x,i)$.
Examples of the construction of $\lambda(x,i)$ occur below as part of Examples~\ref{can ex 1} and~\ref{can ex 2}.
\begin{prop}\label{min below with inv}
The permutation $\lambda(x,i)$ is the unique minimal element of the set $\set{y\le x:(x_i,x_{i+1})\in\operatorname{inv}(y)}$.
\end{prop}
\begin{proof}
The inversion set of $\lambda(x,i)$ consists of all pairs $(b,a)$ with ${x_i\ge b>a\ge x_{i+1}}$, with $b\in\set{x_j:1\le j\le i}$, and with $a\in\set{x_j:i+1\le j\le n}$.
Each of these is an inversion of $x$, so $\lambda(x,i)\le x$.
Given a permutation $y$ with $y\le x$ and $(x_i,x_{i+1})\in\operatorname{inv}(y)$ and such a pair $(b,a)$, we will show that $(b,a)$ is an inversion of $y$.
The elements $b$, $a$, $x_i$ and $x_{i+1}$ occur in the order $\cdots b\cdots x_ix_{i+1}\cdots a\cdots$ in $x$.
Also $(x_i,x_{i+1})\in\operatorname{inv}(y)$, or in other words $x_i$ precedes $x_{i+1}$ in $y$.
If $(b,a)\not\in\operatorname{inv}(y)$, or in other words if $a$ precedes $b$ in $y$, then necessarily either $a$ precedes $x_{i+1}$ or $b$ follows $x_i$ in $y$, or both.
Thus either $(a,x_{i+1})$ or $(x_i,b)$ is in $\operatorname{inv}(y)$, contradicting the fact that $y\le x$.
We have shown that $\operatorname{inv}(\lambda(x,i))\subseteq y$, and thus ${\lambda(x,i)\le y}$.
\end{proof}
The next theorem says in particular that the canonical joinands of a permutation are in bijection with the descents of the permutation.
\begin{theorem}\label{perm can join}
The canonical join representation of a permutation $x$ is
\[x=\Join\set{\lambda(x,i):x_i>x_{i+1}}.\]
\end{theorem}
\begin{proof}
We first show that $\Join\set{\lambda(x,i):x_i>x_{i+1}}=x$.
Proposition~\ref{min below with inv} implies in particular that $\lambda(x,i)\le x$ for each $i$ with $x_i>x_{i+1}$, so ${\Join\set{\lambda(x,i):x_i>x_{i+1}}\le x}$.
If $\Join\set{\lambda(x,i):x_i>x_{i+1}}<x$, then some $y\lessdot x$ has $y\ge\Join\set{\lambda(x,i):x_i>x_{i+1}}$.
The permutation $y$ is obtained from $x$ by swapping $x_i$ and $x_{i+1}$ for some $i$ such that $x_i>x_{i+1}$.
But then $(x_i,x_{i+1})$ is not an inversion of $y$, while it is an inversion of $\lambda(x,i)$, and therefore $y\not\ge\lambda(x,i)$, and this is a contradiction.
We conclude that $\Join\set{\lambda(x,i):x_i>x_{i+1}}=x$.
We next show that $\set{\lambda(x,i):x_i>x_{i+1}}$ is an antichain.
Suppose $i$ and $j$ are distinct indices with $x_i>x_{i+1}$ and $x_j>x_{j+1}$.
The inversion set of $\lambda(x,i)$ is described in the proof of Proposition~\ref{min below with inv}.
In particular, if $(x_j,x_{j+1})$ is an inversion of $\lambda(x,i)$, then $j\le i$ and $i+1\le j+1$, which is impossible because $i\neq j$.
Therefore $(x_j,x_{j+1})$ is not an inversion of $\lambda(x,i)$.
Since $(x_j,x_{j+1})$ \emph{is} an inversion of $\lambda(x,j)$, we see that $\lambda(x,j)\not\le\lambda(x,i)$.
Thus $\set{\lambda(x,i):x_i>x_{i+1}}$ is an antichain.
Finally, we show that $\set{\lambda(x,i):x_i>x_{i+1}}\ll T$ whenever $x=\Join T$.
Suppose to the contrary that for some $i$ with $x_i>x_{i+1}$, there is no element $t\in T$ with $\lambda(x,i)\le t$.
Then Proposition~\ref{min below with inv} implies that there is no element $t\in T$ with $(x_i,x_{i-1})\in\operatorname{inv}(t)$.
Let $y$ be the permutation obtained from $x$ by swapping $x_i$ and $x_{i+1}$.
Then $\operatorname{inv}(y)=\operatorname{inv}(x)\setminus\set{(x_i,x_{i+1})}$.
But since the weak order is containment of inversion sets and since $x$ is an upper bound for $T$, we see that $y$ is also an upper bound for $T$, contradicting the supposition that $x=\Join T$.
\end{proof}
\begin{example}\label{can ex 1}
In Figure~\ref{weak ex}, we see directly that the canonical joinands of the permutation $3421$ are $2134$ and $1342$, in agreement with Theorem~\ref{perm can join}.
\end{example}
\begin{example}\label{can ex 2}
Let $x$ be the permutation $157842936\in S_9$.
Then $x$ has descents $8>4$ and $4>2$ and $9>3$.
The canonical joinands of $x$ are $\lambda(x,4)=123578469$, $\lambda(x,5)=142356789$ and $\lambda(x,7)=124578936$.
\end{example}
Theorem~\ref{perm can join} describes the join representation of a given permutation, but does not answer the following natural question.
\begin{question}\label{can q}
Which sets of join-irreducible permutations are canonical join representations?
\end{question}
In the next section, we answer Question~\ref{can q} using noncrossing arc diagrams.
\section{Canonical join representations and noncrossing arc diagrams}\label{proof sec}
In this section, we give a bijection between permutations and noncrossing arc diagrams by showing that noncrossing arc diagrams are a combinatorial model for canonical join representations.
A permutation is join-irreducible if and only if it has exactly one descent.
Suppose $x_1\cdots x_n$ is a join-irreducible permutation with descent $x_i>x_{i+1}$.
The unique permutation covered by $x$ is obtained by swapping the adjacent entries $x_i$ and $x_{i+1}$.
In particular, $x$ itself is the unique minimal element of $\set{y\le x:(x_i,x_{i+1})\in\operatorname{inv}(y)}$, so Proposition~\ref{min below with inv} says that $x=\lambda(x,i)$.
Thus, given that $x$ is join-irreducible, it is determined uniquely by the values $x_i$ and $x_{i+1}$ forming its unique descent and by the set of values $\set{x_j:j<i,\,x_{i+1}<x_j<x_i}$.
(To specify $x$, we don't need to specify $i$ explicitly; the two values and the set are enough.)
On the other hand, an arc satisfying (A) is determined by its endpoints and by the set of points to its left.
Thus there is a bijection between join-irreducible permutations and arcs satisfying (A).
The bijection sends a join-irreducible permutation $x$ to the arc connecting $a$ and $b$, where $b>a$ is the unique descent of $x$, and for each $c$ with $a<c<b$, having $c$ to the left of the arc if and only if $c$ occurs in $x$ to the left of the descent $ba$.
We extend this bijection to a map from permutations to sets of arcs.
Specifically, let $\delta$ be the map taking a permutation to the set of arcs corresponding to the elements of its canonical join representation.
The following theorem constitutes a bijective proof of Theorem~\ref{diagrams perm}.
\begin{theorem}\label{main}
The map $\delta$ is a bijection from $S_n$ to the set of noncrossing arc diagrams on $n$ points.
\end{theorem}
As a first step in proving the theorem, we recast the map $\delta$ in way that makes it clear that it maps permutations to noncrossing arc diagrams.
Given a permutation $x=x_1,\ldots,x_n$, write each entry $x_i$ at the point $(i,x_i)$ in the plane.
For each descent $x_i>x_{i+1}$, draw a line from $x_i$ to $x_{i+1}$ as illustrated in the left picture of Figure~\ref{map to ncs} for the permutation $157842936$.
(Cf. Example~\ref{can ex 2}.)
\begin{figure}
\scalebox{0.9}{\includegraphics{map1}}\qquad\qquad\quad\scalebox{0.9}{\includegraphics{map2}}
\caption{The map from permutations to noncrossing arc diagrams}
\label{map to ncs}
\end{figure}
Then move all of the numbers into a single vertical line, allowing the lines connecting descents to curve but not to pass through any of the numbers.
These lines become the arcs in a noncrossing arc diagram.
By Theorem~\ref{perm can join}, we see that these arcs correspond to the elements of the canonical join representation of~$x$, so the noncrossing arc diagram obtained is exactly~$\delta(x)$.
It will be convenient to prove Theorem~\ref{main} together with the following result, which was promised in the introduction.
\begin{prop}\label{pairwise works}
Given any collection of pairwise compatible arcs, there is a noncrossing arc diagram whose arcs are combinatorially equivalent to the given arcs.
\end{prop}
\begin{proof}[Proof of Theorem~\ref{main} and Proposition~\ref{pairwise works}]
Since a permutation is uniquely determined by its canonical join representation, the map $\delta$ is injective.
Let $\mathcal{E}$ be some collection of pairwise compatible arcs, each satisfying condition (A) of the definition of noncrossing arc diagrams.
We do not yet know that we can draw all the arcs of $\mathcal{E}$ together in such a way that condition (C1) holds.
However, we know that (A) holds for each arc in $\mathcal{E}$ and that (C1) and (C2) hold for each pair.
Consider the graph $\mathcal{G}$ defined on the given $n$ points with edges given by $\mathcal{E}$.
By (C2), each connected component of $\mathcal{G}$ is either an isolated point or a sequence $i_1>\cdots>i_k$ such that each $i_j$ and $i_{j+1}$ are connected by an arc in $\mathcal{E}$.
In Section~\ref{diagrams sec}, in connection with the definition of compatibility of arcs, we described combinatorial conditions that would require one arc to be drawn to the left or right of another.
Say a component $C_1$ of $\mathcal{G}$ is \newword{left of} another component $C_2$ (or equivalently that $C_2$ is \newword{right of} $C_1$) if there exist arcs $\alpha_1$ in $C_1$ and $\alpha_2$ in $C_2$ such that $\alpha_1$ must be drawn to the left of $\alpha_2$.
We claim that the relation ``is left of'' is acyclic on components of $\mathcal{G}$.
There are no 1-cycles in the relation.
A 2-cycle in the relation is a pair $C_1,C_2$ of components such that $C_1$ is both right and left of $C_2$.
If $C_2$ is left of $C_1$, then there exists a point $p$ that is left of (or an endpoint of) an arc in $C_1$ and is right of (or an endpoint of) an arc in $C_2$, with $p$ not an endpoint of both arcs.
At $p$, there is an arc $\alpha_1$ of $C_1$ that must be right of some arc $\alpha_2$ in $C_2$.
As one moves vertically upwards from $p$, there may be endpoints of arcs in $C_1$ or $C_2$.
As we pass through such an endpoint (say in $C_1$) and pass to a new arc in $C_1$, condition (C2) says that we don't also pass through an endpoint of $C_2$.
We pass to a new arc $\alpha_1'$ in $C_1$ which also must be right of $\alpha_2$.
Thus, continuing upwards and passing endpoints of arcs in $C_1$ or $C_2$, we find that we must continue to draw $C_1$ right of $C_2$.
After also making the same argument moving downward from $p$, we see that \emph{all} of $C_1$ can be drawn to the right of $C_2$.
In particular, $C_1$ is not left of $C_2$, so $C_1,C_2$ is not a 2-cycle.
Now suppose that $k\ge 3$ and $C_1,\ldots,C_k$ are components of $\mathcal{G}$ with the property that $C_{i+1}$ is left of $C_i$ for each $i=1,\ldots k$.
We will index cyclically throughout the argument, so that in particular $C_1$ is left of $C_k$.
Each $C_i$ has some lowest vertex $a_i$ and some highest vertex $b_i$.
First, suppose the interval $[a_i,b_i]$ is contained in the interval $[a_{i+1},b_{i+1}]$ for some $i$.
Arguing as in the proof that $2$-cycles don't exist, we see that all of $C_i$ must be drawn to the right of $C_{i+1}$ to satisfy (C2).
There is a point $p$ that is left of (or an endpoint of) some arc in $C_{i-1}$ and right of (or an endpoint of) some arc in $C_i$.
The point $p$ is also right of $C_{i+1}$, and we conclude that $C_{i+1}$ is left of $C_{i-1}$.
Thus we obtain a $(k-1)$-cycle by removing $C_i$ from $C_1,\ldots,C_k$.
If instead $[a_i,b_i]$ contains $[a_{i+1},b_{i+1}]$, the we similarly make a $(k-1)$-cycle.
Now suppose that for all $i$, there is no containment relation between $[a_i,b_i]$ and $[a_{i+1},b_{i+1}]$.
Since $C_1,\ldots,C_k$ is a cycle, there exists $i$ with ${a_{i-1}<a_i<b_{i-1}<b_i}$ and ${a_{i+1}<a_i<b_{i+1}<b_i}$.
Then the point $a_i$ is to the left of some arc of $C_{i-1}$ and to the right of some arc of $C_{i+1}$.
Thus again we obtain a $(k-1)$-cycle by removing $C_i$ from $C_1,\ldots,C_k$.
We have shown that, for $k\ge3$, if there exists a $k$-cycle in the ``is left of'' relation, then there exists a $(k-1)$-cycle.
Since $1$-cycles and $2$-cycles don't exist, this completes the proof of the claim that the relation is acyclic.
We now use the claim to recursively construct a permutation $x=x_1\cdots x_n$ with the following properties:
\begin{enumerate}
\item[(i)]If $i>j$, then $i$ and $j$ form a descent in $x$ if and only if $i$ and $j$ are connected by an arc in $\mathcal{E}$.
\item[(ii)]If $i>j>k$ and $i$ and $k$ are connected by an arc in $\mathcal{E}$, then $j$ is to the left of $i$ in $x$ if and only if $j$ is to the left of the arc connecting $i$ to $k$.
(Equivalently, $j$ is right of $k$ in $x$ if and only if $j$ is right of the arc.)
\end{enumerate}
The claim implies that $\mathcal{G}$ has at least one \newword{left component}, meaning a component that is not right of any other component of $\mathcal{G}$.
Since each left component has nothing to its left, the left components can be totally ordered from smaller-valued endpoints to higher-valued ones.
Take the smallest-valued left component, delete it from $\mathcal{G}$ and write its labels in decreasing order.
(When we delete the component, we keep the original labels on the remaining points of $\mathcal{G}$.)
By the claim, if the remaining graph is nonempty, then it has at least one left component, and in particular may have additional left components that were not left components in the original diagram.
We again take the smallest left component, delete it, and write its labels in decreasing order.
When all of the points of the diagram have been deleted, the result is a permutation.
The output of the process is a permutation that satisfies condition (ii) above by construction.
Because the arcs in $\mathcal{E}$ satisfy (C2) pairwise, each arc in the diagram becomes a descent in the permutation.
It is also easy to see that there are no other descents:
If there were another descent, that would mean that at some step the process deletes one left component $C_1$, and then another left component $C_2$ whose highest point is lower than the lowest point of $C_1$.
But then $C_2$ was already a left component before $C_1$ was deleted, so $C_2$ should have been deleted before $C_1$.
By this contradiction, we see that the output permutation $x$ satisfies (i).
Conditions (i) and (ii) imply that $\delta(x)$ is a noncrossing arc diagram whose arcs are exactly $\mathcal{E}$.
In particular, we have proved Proposition~\ref{pairwise works}.
We have also shown that $\delta$ is surjective, thus completing the proof that it is a bijection.
\end{proof}
The inverse to $\delta$ is given by the recursive process described in the proof above, as exemplified in Figure~\ref{inversefig}.
At each step, the left components of the remaining diagram are shown in red (or gray if the figure is not viewed in color).
\begin{figure}
\begin{tabular}{|l|l|l|l|l|l|l|}
\hline&&&&&&\\[-8pt]
Step&Start&1&2&3&4&5\\[2 pt]\hline&&&&&&\\[-8pt]
Permutation so far&&4&46&46731&4673152&46731528\\[2 pt]\hline&&&&&&\\[-8pt]
\raisebox{59pt}{Diagram remaining}&\scalebox{0.85}{\includegraphics{inverse0}}&\scalebox{0.85}{\includegraphics{inverse1}}&\scalebox{0.85}{\includegraphics{inverse2}}&\scalebox{0.85}{\includegraphics{inverse3}}&\scalebox{0.85}{\includegraphics{inverse4}}&\\[2 pt]\hline
\end{tabular}
\caption{The map from noncrossing arc diagrams to permutations}
\label{inversefig}
\end{figure}
\begin{remark}\label{foreshadowing}
Noncrossing arc diagrams and the map $\delta$ are foreshadowed in work~\cite{shardint} on \newword{shard intersections} and particularly in work of Bancroft~\cite{Bancroft} and Petersen~\cite{Petersen} on shard intersections in type A.
See in particular \cite[Theorem~3.6]{shardint}, \cite[Proposition~4.7]{shardint}, \cite[Section~3]{Bancroft}, and \cite[Section~2.1]{Petersen}.
\end{remark}
The noncrossing arc diagrams are the faces of a simplicial complex whose vertices are the arcs.
The face corresponding to a noncrossing arc diagram is the set of arcs appearing in the diagram.
The following is an immediate corollary of Theorem~\ref{main}.
\begin{cor}\label{main can join cplx}
The simplicial complex of noncrossing arc diagrams on $n$ points is isomorphic to the canonical join complex of the weak order on $S_n$.
The isomorphism is induced by the map from arcs to join-irreducible elements.
\end{cor}
Figure~\ref{arc complex} shows the canonical join complex of the weak order on $S_4$, in the form of the complex of noncrossing arc diagrams.
\begin{figure}
\includegraphics{complex}
\caption{The canonical join complex of $S_4$}
\label{arc complex}
\end{figure}
Theorem~\ref{main} and Proposition~\ref{pairwise works} also lead to an answer to Question~\ref{can q}.
In the following corollary, the equivalence of (i) and (ii) is immediate by Theorem~\ref{main} and the equivalence of these with (iii) and (iv) follows by Proposition~\ref{pairwise works}.
\begin{cor}\label{can a}
Suppose $J$ is a set of join-irreducible elements of $S_n$ and suppose $\mathcal{E}$ is the corresponding collection of arcs.
The following are equivalent.
\begin{enumerate}
\item[\textnormal{(i)}] $J$ is the canonical join representation of a permutation.
\item[\textnormal{(ii)}] There is a noncrossing arc diagram whose arcs are combinatorially equivalent to the arcs in $\mathcal{E}$.
\item[\textnormal{(iii)}] The arcs in $\mathcal{E}$ are pairwise compatible.
\item[\textnormal{(iv)}] Each $2$-element subset of $J$ is the canonical join representation of a permutation.
\end{enumerate}
\end{cor}
A simplicial complex is called \newword{flag} it it is the clique complex of its $1$-skeleton.
Equivalently, it is flag if, for every subset $S$ of the vertices not forming a face, there is a pair of distinct elements of $S$ not forming an edge in the complex.
The following corollary is immediate by Corollary~\ref{can a}.
\begin{cor}\label{arcs flag}
The canonical join complex of the weak order on permutations is flag.
\end{cor}
Theorem~\ref{main} also immediately implies some additional counting results, and combines with known results to prove others.
Let $\genfrac{\langle}{\rangle}{0pt}{}{n}{k}$ denote the Eulerian number, the number of permutations of $\set{1,\ldots,n}$ with exactly $k$ descents.
By Theorem~\ref{perm can join}, the descents of a permutation are in bijection with the join-irreducible permutations in its canonical join representation.
The latter are in bijection with the arcs in the corresponding noncrossing arc diagram, so we have the following theorem.
\begin{theorem}\label{diagrams eulerian}
The number of noncrossing arc diagrams on $n$ points with exactly $k$ arcs is $\genfrac{\langle}{\rangle}{0pt}{}{n}{k}$.
\end{theorem}
Say a noncrossing arc diagram is a \newword{matching} if all of its arcs are disjoint, even at their endpoints.
In this case, the diagram is a planar representation of a matching in the usual graph-theoretic sense.
A noncrossing arc diagram is a matching if and only if the corresponding permutation $x$ has no three consecutive entries $x_ix_{i+1}x_{i+2}$ with $x_i>x_{i+1}>x_{i+2}$.
Such permutations are said to \newword{avoid the consecutive pattern $321$}.
The following theorem is an immediate consequences of Theorem~\ref{main}.
\begin{theorem}\label{diagrams matchings}
The noncrossing arc diagrams on $n$ points that are matchings are in bijection with permutations avoiding the consecutive pattern $321$.
\end{theorem}
The exponential generating function for permutations avoiding the consecutive pattern $321$ (and thus for noncrossing arc diagrams that are matchings) is determined in~\cite[Theorem~4.1]{ElizaldeNoy}.
Say a noncrossing arc diagram is a \newword{perfect matching} if it is a matching and each point is the endpoint of an arc.
An \newword{alternating permutation}\footnote{Conventions vary: Sometimes the term alternating permutation refers instead to permutations satisfying $x_1<x_2>x_3<x_4>\cdots >x_{2n-1}<x_{2n}$, and sometimes the term refers to permutations satisfying either of the two conditions.} in $S_{2n}$ is a permutation $x$ with $x_1>x_2<x_3>x_4<\cdots <x_{2n-1}>x_{2n}$.
Alternating permutations in $S_{2n}$ are characterized by avoiding the consecutive pattern $321$ and having exactly $n$ descents.
Thus $\delta$ restricts to the bijection described in the following theorem.
\begin{theorem}\label{diagrams perfect}
Alternating permutations in $S_{2n}$ are in bijection with noncrossing arc diagrams on $2n$ points that are perfect matchings.
\end{theorem}
The exponential generating function for alternating permutations (and thus for noncrossing arc diagrams that are perfect matchings) is $\sec x$.
Left noncrossing perfect matchings in the sense of this paper are also well-known, usually under the name ``noncrossing matchings.''
Restricting $\delta$, these are in bijection with $231$-avoiding alternating permutations.
From \cite[Theorem~2.2]{Mansour} (for the permutations) or from \cite[Exercise~6.19(o)]{EC1} (for the matchings), we obtain the following enumeration:
The number of left noncrossing arc diagrams on $2n$ points that are perfect matchings is the Catalan number $C_n=\frac1{n+1}\binom{2n}n$.
\section{Noncrossing arc diagrams and lattice quotients of the weak order}\label{quot sec}
In this section, we discuss how restricted classes of noncrossing arc diagrams arise from lattice quotients of the weak order modulo lattice congruences.
We begin by briefly reviewing some background on lattice congruences.
We then describe how noncrossing arc diagrams are a convenient combinatorial model for lattice quotients of the weak order.
Finally, we give some examples of restricted classes of noncrossing arc diagrams arising from quotients.
A \newword{congruence} on a lattice $L$ is an equivalence relation on $L$ that respects the meet and join operations.
That is, if $x_1\equiv y_1$ and $x_2\equiv y_2$, then $x_1\wedge x_2\equiv y_1\wedge y_2$ and $x_1\vee x_2\equiv y_1\vee y_2$.
The property of respecting meets and joins is equivalent, \emph{for finite lattices}, to the following three conditions:
First, equivalence classes are intervals in the lattice.
Second, the map $\pi_\downarrow^\Theta$ taking each element to the bottom element of its equivalence class is order-preserving.
Third, the map $\pi^\uparrow_\Theta$ taking each element to the top element of its equivalence class is order-preserving.
The \newword{quotient} of $L$ modulo a congruence $\Theta$ is the lattice $L/\Theta$ whose elements are the congruence classes with the meet or join of classes defined by taking the meet or join of representatives.
That is, if $C$ and $D$ are congruence classes with $x\in C$ and $y\in D$, then $C\wedge D$ is the class of $x\wedge y$ and $C\vee D$ is the class of $x\vee y$.
\emph{For finite lattices}, the quotient is isomorphic as a poset to the subposet $\pi_\downarrow^\Theta(L)$ of $L$ induced by the elements that are the bottom elements of their congruence classes.
(The subposet $\pi_\downarrow^\Theta(L)$ is always a join-sublattice of $L$ but can fail to be a sublattice of~$L$.)
This way of thinking about the quotient suggests that we think of ``contracting'' each congruence class onto its bottom element.
An element of $L$ is \newword{contracted} by $\Theta$ if it is congruent, modulo $\Theta$, to some element below it.
Thus an element is \newword{uncontracted} if and only if it is at the bottom of its congruence class, and so $\pi_\downarrow^\Theta(L)$ is the subposet of $L$ induced by uncontracted elements.
In a lattice $L$ where each element has a canonical join representation, an element is contracted by $\Theta$ if and only if one or more of its canonical joinands is contracted by $\Theta$.
In particular, the join-irreducible elements of $\pi_\downarrow^\Theta(L)$ are exactly the join-irreducible elements of $L$ not contracted by $\Theta$.
Furthermore, the canonical join-representation of an element of the quotient $\pi_\downarrow^\Theta(L)$ coincides with its canonical join-representation in $L$.
Thus Theorem~\ref{main} implies the following theorem.
\begin{theorem}\label{quot diagram}
Given a lattice congruence $\Theta$ on the weak order on $S_n$, the elements of the quotient lattice $\pi_\downarrow^\Theta(S_n)$ are in bijection with the noncrossing arc diagrams on $n$ points consisting only of arcs corresponding to join-irreducible elements not contracted by $\Theta$.
The bijection maps an element of $\pi_\downarrow^\Theta(S_n)$ to the set of arcs corresponding to its canonical join-representation (in $\pi_\downarrow^\Theta(S_n)$ or in $S_n$).
\end{theorem}
The canonical join complex of $\pi_\downarrow^\Theta(L)$ is isomorphic to the subcomplex of the canonical join complex of $L$ induced by the vertices (join-irreducible elements of $L$) not contracted by $\Theta$.
Thus we have the following additional corollaries.
(Compare Corollaries~\ref{main can join cplx} and~\ref{arcs flag}.)
\begin{cor}\label{quot can join cplx}
Given a lattice congruence $\Theta$ on the weak order on $S_n$, the canonical join complex of $\pi_\downarrow^\Theta(S_n)$ is isomorphic to the simplicial complex of noncrossing arc diagrams on $n$ points using only arcs corresponding to join-irreducible elements not contracted by $\Theta$.
The isomorphism is induced by the map from join-irreducible elements to arcs.
\end{cor}
\begin{cor}\label{arcs quot flag}
For any lattice congruence $\Theta$ on the weak order on $S_n$, the canonical join complex of $S_n/\Theta$ is flag.
\end{cor}
The fact that an element is contracted if and only if one or more of its canonical joinands is contracted implies in particular that the congruence is determined uniquely by which join-irreducible elements it contracts.
As one might expect, the join-irreducible elements cannot be contracted independently.
Instead, there is a pre-order on join-irreducible elements such that a set of join-irreducible elements is contracted by some congruence if and only if that set is closed under ``going down'' in the preorder.
We refer to this preorder as \newword{forcing} and describe it in words rather than notation.
A join-irreducible element $j_1$ is above $j_2$ in the forcing order if every congruence contracting $j_1$ also contracts $j_2$.
In this case, we say that $j_1$ \newword{forces} $j_2$.
The forcing preorder on join-irreducible elements of the weak order on $S_n$ was worked out in \cite[Section~8]{congruence}.
In particular, in the weak order on $S_n$, the forcing preorder was shown to be a partial order (i.e.\ it has no directed cycles).
The forcing order on join-irreducible permutations has a nice description in terms of arcs satisfying condition (A).
We will phrase the description by saying that one arc forces another arc, meaning that the forcing relation holds on the corresponding join-irreducible permutations.
Suppose $\alpha_1$ and $\alpha_2$ are arcs satisfying (A), with $\alpha_1$ connecting points $p_1$ and $q_1$ with $p_1<q_1$ and with $\alpha_2$ connecting $p_2$ and $q_2$ with $p_2<q_2$.
We say $\alpha_1$ is a \newword{subarc} of $\alpha_2$ if
\begin{enumerate}
\item[(i) ]$p_2\le p_1<q_1\le q_2$, and
\item[(ii) ]The set of points left of $\alpha_1$ equals the set of points in $\set{p_1+1,\ldots,q_1-1}$ left of $\alpha_2$.
\end{enumerate}
Less formally, to construct a subarc of an arc $\alpha$, we choose two distinct horizontal lines, each passing through one of the $n$ points and each intersecting $\alpha$, possibly at an endpoint of $\alpha$.
We construct the subarc by cutting $\alpha$ along the two lines and retaining the middle portion of $\alpha$.
Each endpoint of this middle section is attached to one of the $n$ points, specifically, the point at the same height.
(Possibly the endpoint of the middle section is already one of the $n$ points, if the line cuts $\alpha$ at an endpoint.)
This process is illustrated in Figure~\ref{subarcfig}.
\begin{figure}
\begin{tabular}{ccccc}
\scalebox{0.8}{\includegraphics{subarc0}}&
\scalebox{0.8}{\includegraphics{subarc1}}&
\scalebox{0.8}{\includegraphics{subarc2}}&
\scalebox{0.8}{\includegraphics{subarc3}}&
\scalebox{0.8}{\includegraphics{subarc4}}\\
\end{tabular}
\caption{Constructing a subarc}
\label{subarcfig}
\end{figure}
In \cite[Section~8]{congruence}, the join-irreducible permutations are encoded as subsets as explained in \cite[Section~5]{congruence}.
Translating this encoding into the language of arcs and subarcs, either \cite[Theorem~8.1]{congruence} or \cite[Theorem~8.2]{congruence} immediately implies the following theorem:
\begin{theorem}\label{arc forcing}
An arc $\alpha_1$ forces an arc $\alpha_2$ if and only if $\alpha_1$ is a subarc of $\alpha_2$.
\end{theorem}
Figure~\ref{forcingfig} shows the forcing order on join-irreducible permutations in $S_4$, represented by arcs.
\begin{figure}
\includegraphics{forcing}
\caption{The forcing order on arcs for $n=4$}
\label{forcingfig}
\end{figure}
Theorems~\ref{quot diagram} and~\ref{arc forcing} let us understand quotients of the weak order on $S_n$ entirely in terms of noncrossing arc diagrams.
(It is not immediately apparent how to realize the partial order/lattice structure on the quotient in terms of noncrossing arc diagrams, so we confine ourselves to statements about the quotients as sets.)
The following corollary is immediate by Theorem~\ref{arc forcing} (for the first assertion), then Theorem~\ref{quot diagram} (for the second and third assertions).
\begin{cor}\label{closed subarcs}
A set $U$ of arcs corresponds to the set of \emph{un}contracted join-irreducible permutations of some congruence $\Theta$ if and only if $U$ is closed under passing to subarcs.
In this case, the map $\delta$ restricts to a bijection from permutations not contracted by $\Theta$ to arc diagrams consisting only of arcs in $U$.
For each $k$, the map $\delta$ further restricts to a bijection between uncontracted permutations with exactly $k$ descents and arc diagrams consisting of exactly $k$ arcs, all of which are in~$U$.
\end{cor}
We now give an explicit description of the uncontracted permutations.
Suppose $a$ and $b$ are integers with $1\leq a<b\leq n$ and suppose $R\subseteq\set{(a+1),\ldots,(b-1)}$.
Write $L=\set{(a+1),\ldots,(b-1)}\setminus R$.
A permutation $x$ \newword{has a $(b,a,R)$-pattern} if $(x_i,x_{i+1})$ is a descent of $x$, if $x_i\geq b$ and $x_{i+1}\leq a$, and if all of the elements of $L$ appear in $x$ \emph{to the left of} $x_i,x_{i+1}$ and all of the elements of $R$ appear in $x$ \emph{to the right of} $x_i,x_{i+1}$.
If $x$ has no $(b,a,R)$-pattern, then $x$ \newword{avoids $(b,a,R)$}.
The triple $(b,a,R)$ precisely specifies an arc $\alpha(b,a,R)$ connecting $a$ and $b$ and having the points in $R$ on its right while having the points in $L$ on its left.
The same information precisely specifies a join-irreducible permutation $\lambda(b,a,R)$ consisting of the entries $1,\ldots,(a-1)$, then $L$ in increasing order, then $b$, then $a$, then $R$ in increasing order, and finally $(b+1),\ldots,n$.
Our bijection between join-irreducible permutations and arcs sends $\lambda(b,a,R)$ to $\alpha(b,a,R)$.
\begin{cor}\label{perm quot}
Let $\Theta$ be the smallest congruence contracting the join-irreducible permutations in a given set $\set{\lambda(b_i,a_i,R_i):i\in I}$.
Then the permutations not contracted by $\Theta$ are exactly the permutations that avoid $(b_i,a_i,R_i)$ for every $i\in I$.
\end{cor}
\begin{proof}
By Corollary~\ref{closed subarcs}, a permutation $x$ is uncontracted by $\Theta$ if and only if no arc in $\delta(x)$ has an arc $\alpha(b_i,a_i,R_i)$ as a subarc.
By the definition of $\delta$, this description of the uncontracted permutations is a criterion on the canonical join-representations of the permutations.
Theorem~\ref{perm can join} translates the criterion into the requirement of avoiding $(b_i,a_i,R_i)$ for every $i\in I$.
\end{proof}
The results of \cite{con_app} imply a useful enumerative statement.
The Hasse diagram of the weak order on permutations is dual to the simplicial fan $\mathcal{F}$ defined by the Coxeter arrangement of type A.
Given a congruence $\Theta$ on the weak order, for each congruence class, one can ``glue'' together the cones corresponding to the elements of the congruence class.
In \cite[Theorem~1.1]{con_app}, it is shown, among other things, that for each congruence class, the result of the gluing is a convex cone, that these glued cones form a fan $\mathcal{F}_\Theta$, and that any linear extension of the quotient lattice defines a shelling order on $\mathcal{F}_\Theta$.
When $\mathcal{F}_\Theta$ is simplicial, its $h$-vector (i.e.\ the $h$-vector of the corresponding simplicial complex) counts permutations not contracted by $\Theta$ according to their number of descents.
The fan $\mathcal{F}_\Theta$ is simplicial if and only if the Hasse diagram of the quotient is a regular graph.
By Corollary~\ref{closed subarcs}, we have the following corollary.
\begin{cor}\label{simp h}
Suppose $\Theta$ is a congruence on the weak order on permutations and suppose $U$ is the set of arcs corresponding to join-irreducible permutations not contracted by $\Theta$.
If the fan $\mathcal{F}_\Theta$ is simplicial, then the entry $h_k$ in the $h$-vector of $\mathcal{F}_\Theta$ counts noncrossing arc diagrams consisting of exactly $k$ arcs, all of which are in $U$.
\end{cor}
We conclude with some examples.
\begin{example}[Permutations with restricted size of descent]
The \newword{length} of an arc is $1$ plus the number of points the arc passes either left or right of.
If the points are evenly spaced at unit distance, then the length of an arc is the distance between its endpoints.
Thus the arcs in a noncrossing arc diagram on $n$ points have lengths $1$ through $n-1$.
Fixing some $k\ge 1$, the set of arcs of length less than~$k$ is closed under passing to subarcs, so there is a lattice quotient of the weak order on $S_n$ corresponding to noncrossing arc diagrams with arcs of length less than~$k$.
The corresponding congruence is the smallest congruence that contracts all join-irreducible permutations corresponding to arcs of length~$k$.
The uncontracted permutations $x_1x_2\cdots x_n$ are characterized by the requirement that $x_i-x_{i+1}<k$ for $i=1,\ldots n-1$.
These permutations are also easily counted by induction on~$n$, and we see that the number of noncrossing arc diagrams on $n$ points with arcs of length less than~$k$ is $\prod_{i=1}^n\min(i,k)$.
This is $n!$ for $n\le k$ and $k!k^{n-k}$ for $k\le n$.
\end{example}
\begin{example}[The Tamari lattice and Cambrian lattices of type A]
The set of left arcs is closed under passing to subarcs.
Thus the permutations corresponding to left noncrossing arc diagrams (as defined in the introduction) are a lattice quotient $S_n/\Theta$ of the weak order on $S_n$.
Then $\Theta$ is the smallest congruence contracting all join-irreducible permutations corresponding to right arcs of length $2$.
Thus by Corollary~\ref{perm quot}, permutations not contracted by $\Theta$ are exactly the permutations that avoid $231$ in the usual sense.
We see that $S_n/\Theta$ is the Tamari lattice.
More generally, arbitrarily designate each of the $n$ points either as a right point or a left point.
Consider the set $U$ of arcs that do not pass to the right of any right point and do not pass to the left of any left point.
This set is closed under passing to subarcs.
Let $\Theta$ be the congruence that leaves uncontracted exactly the join-irreducible permutations corresponding to arcs in $U$.
Then $\Theta$ is the smallest congruence contracting the join-irreducible permutations corresponding to arcs of length $2$ that pass right of a right point or left of a left point.
The quotient $S_n/\Theta\cong \pi_\downarrow^\Theta(S_n)$ is a \newword{Cambrian lattice} of type~A.
The subposet $\pi_\downarrow^\Theta(S_n)$ has the special property that it is a sublattice of $S_n$.
The fan $\mathcal{F}_\Theta$ is simplicial, and in fact is the normal fan to an associahedron.
Corollary~\ref{simp h} reflects the well-known fact that the $h$-vector of the associahedron is given by the Narayana numbers.
For more information on the Tamari lattice and Cambrian lattice of type~A, see \cite[Sections~5--6]{cambrian}.
\end{example}
\begin{example}[Twisted Baxter permutations/diagrams with left and right arcs]
The set consisting of all left arcs and all right arcs is closed under passing to subarcs.
Thus the noncrossing arc diagrams having only left arcs and right arcs constitute a lattice quotient of the weak order.
This quotient is mentioned in \cite[Section~10]{con_app} and studied extensively in \cite{rectangle}.
Let $\Theta$ be the corresponding congruence.
The left and right arcs are exactly the arcs with no inflections (as defined in Section~\ref{diagrams sec}), so $\Theta$ is the smallest congruence contracting all join-irreducible permutations whose corresponding arcs have length $3$ and one inflection point.
For $n\ge 4$, the subposet $\pi_\downarrow^\Theta(S_n)$ consisting of uncontracted permutations is not a sublattice of the weak order, and the corresponding fan is not simplicial.
(Both of these facts are easily verified for $n=4$ with the help of Figure~\ref{weak ex}.)
Corollary~\ref{perm quot} implies that $\pi_\downarrow^\Theta(S_n)$ consist of the permutations having no subsequence $x_i\cdots x_jx_{j+1}\cdots x_k$ with $x_{j+1}<x_i<x_j$ and $x_{j+1}<x_k<x_j$.
These are the \newword{twisted Baxter permutations} of \cite[Section~10]{con_app} and \cite{rectangle}, which are known to be in bijection with the \newword{Baxter permutations} of \cite{CGHK}.
(This is an unpublished result of Julian West. For a published proof, see \cite[Theorem~8.2]{rectangle} or combine \cite[Theorem~4.14]{giraudo} and \cite[Proposition~4.15]{giraudo}.
See also \cite[Section~2.4]{Dilks}.)
The Baxter permutations are counted \cite{CGHK} by the Baxter numbers, and Theorem~\ref{diagrams left right} follows.
Theorem~\ref{diagrams left right} is interesting in relation to the \newword{twin binary trees} of Dulucq and Guibert \cite{DGStack}.
These are pairs of planar binary trees that are complementary in a certain sense.
These twin trees can be seen when one cuts a diagonal rectangulation along its diagonal.
In a similar way, a noncrossing arc diagram composed of left arcs and right arcs is a pair of Catalan objects---a left noncrossing arc diagram and a right noncrossing arc diagram---that are compatible in the sense that their union is still a noncrossing arc diagram.
(There are two issues: First, whether the left noncrossing arc diagram and the right noncrossing arc diagram have exactly the same set of arcs connecting adjacent vertices, and second, whether the union satisfies condition~(C2).)
\end{example}
\begin{example}[$2$-clumped permutations/generic rectangulations]
The diagrams whose arcs have at most one inflection point correspond to the \newword{$2$-clumped permutations} of \cite{generic}.
The latter are in bijection \cite[Theorem~4.1]{generic} with generic rectangulations with $n$ rectangles.
Theorem~\ref{diagrams inflection} follows.
More generally, diagrams consisting of arcs with at most $k$ inflection points correspond to the $(k+1)$-clumped permutations.
The $(k+1)$-clumped permutations are not well-studied for $k>1$.
\end{example}
\section*{Acknowledgments}
Thanks to Emily Barnard for helpful comments on several earlier versions of this paper, and in particular for pointing out that Proposition~\ref{pairwise works} needed to be argued.
Thanks also to Vic Reiner and David Speyer for helpful comments.
|
\section{Introduction}
Spontaneous symmetry breaking is a central concept in all of physics~\cite{anderson}. The
quantum collective phenomena induced by zero-point motion of dynamical variables are also
ubiquitous~\cite{lamb,anderson,shender,henley,coleman} and
the dynamics governed by the
quantum fluctuations of the collective degrees of freedom are extremely important for our
understanding of macroscopic emergent phenomena~\cite{caldeira}.
In this paper we demonstrate a different type of spontaneous symmetry breaking, namely, one in which the
macroscopic variables of two interacting many-body systems are locked together through
microscopic zero-point quantum fluctuations. We frame our theory in the context of spinor Bose
gases~\cite{law,koashi,ho,Stamper12}, where quantum fluctuations can be dominating in energetics or dynamics and are highly controllable~\cite{zhou2,turner,barnett,song08,cui}. Specifically, we consider a mixture of two
spinor Bose gases with
interspecies spin exchanges~\cite{shi1,luo,xu,shi2} and show that in the symmetry-breaking ground state,
the spin directors of the
two gases are locked together by the zero-point quantum fluctuations. The spin director is the spin direction modulo $Z_2$ symmetry due to the compensation of the inversion of the spin direction by a $\pi$ phase transformation~\cite{zhou1}. Our results give more motivation to study interspecies spin exchanges in mixtures of different spinor Bose gases~\cite{shi1,luo,xu,shi2}, which are under experimental investigation~\cite{xiong}.
This article is organized as follows.
In Sec.~\ref{hm} we define the many-body Hamiltonian, and discuss the mean-field theory. In Sec.~\ref{qf} we consider quantum fluctuations and find the energies in the Bogoliubov theory. In Sec.~\ref{lock} we investigate the fluctuation-induced locking between the directors of the two species. The fluctuation-induced spin dynamics is discussed in Sec.~\ref{spind}. In Sec.~\ref{trap} we discuss how the locking effect survives a trapping potential or an external magnetic field. Finally, we summarize our investigation in Sec.~\ref{summ}.
\section{Hamiltonian and Mean Field Theory \label{hm} }
The many-body Hamiltonian we apply to study this problem is
\begin{equation}
\mathcal{H}= \sum_{\alpha=a,b}\mathcal{H}_\alpha+\mathcal{H}_{ab},
\end{equation}
$$\begin{array}{rl}
{\cal H}_{\alpha} &= \int d\mathbf{r}
\psi^{\dagger}_{\alpha\mu}h_{\alpha} (\mathbf{r})_{\mu\nu}
\psi_{\alpha\nu} \\ &
+ \frac{1}{2} \int d\mathbf{r}
\psi^{\dagger}_{\alpha\mu}\psi^{\dagger}_{\alpha\rho}
({c}_0^{\alpha}\delta_{\mu\nu}
\delta_{\rho\sigma}+{c}_2^{\alpha}
\mathbf{F}_{\alpha\mu\nu}\cdot\mathbf{F}_{\alpha\rho\sigma})
\psi_{\alpha\sigma}\psi_{\alpha\nu}
\end{array} $$
is the Hamiltonian of species $\alpha$. Here
$h_{\alpha}=-\frac{\hbar^2}{2m_{\alpha}}\nabla^2+V_{\alpha} (\mathbf{r})$ is the spin-independent
single-particle Hamiltonian of each atom of species $\alpha$, $c_0^\alpha$ is the intraspecies density-density interaction strength of
species $\alpha$, $c_2^\alpha$ is the intraspecies spin-exchange interaction strength of
species $\alpha$, and $\psi_{\alpha\mu}$ represents the field
operator for the $\mu$ component of species $\alpha$, with $\alpha=a,b$ and $\mu = -1, 0, 1$ or $x, y ,z$, depending on the basis used. For the case of $\mu=x, y, z$, the $\eta$ component $F^{\eta}_{\mu\nu}$ of
$\bf{F}_{\mu\nu}$ is $- i \epsilon_{\eta \mu\nu}$, where
$\epsilon_{\eta\mu\nu}$ is the Levi-Civit\`{a} antisymmetric tensor~\cite{Snoek04}. In addition,
$${\cal H}_{ab}= \int d\mathbf{r}
\psi^{\dagger}_{a\mu}\psi^{\dagger}_{b\rho}
({c}_0^{ab}\delta_{\mu\nu} \delta_{\rho\sigma}+{c}_2^{ab}
\mathbf{F}_{a\mu\nu}\cdot\mathbf{F}_{b\rho\sigma})
\psi_{b\sigma}\psi_{a\nu}$$
is the interaction between the two species, with $c_0^{ab}$ the interspecies
density-density interaction strength and $c_2^{ab}$ the interspecies spin-exchange
interaction strength. Repeated indices are summed over. We focus on the regime of $c_2^a>0$,
$c_2^b>0$ and $c_2^{ab}>0$ and assume that $c_2^ac_2^b>(c_2^{ab})^2$.
The three-component vector field ${\mathbf \psi}_\alpha$ of the species $\alpha$
can be written as
\begin{equation}
\label{nematic} {\mathbf \psi}_\alpha = \Phi_\alpha\mathbf{n}_\alpha,
\end{equation}
where $\mathbf{n}_\alpha$ is the spin director , which is also a three-component vector, and $$\Phi_\alpha \equiv \sqrt{\rho_\alpha }e^{i\chi_\alpha},$$ where
$$\rho_\alpha=\psi^{\dagger}_{\alpha\mu}\psi_{\alpha\mu}$$ is the number-density operator. In terms of
$\rho_\alpha$ and the spin-density operator $$\textbf{L}_\alpha=
\psi^{\dagger}_{\alpha\mu}\bf{F}_{\alpha\mu\nu}\psi_{\alpha\nu}, $$ the Hamiltonian can be
rewritten as
\begin{equation}
\mathcal{H}=\mathcal{H}_p+\mathcal{H}_s,
\end{equation}
where $$ \begin{array}{rl}
\mathcal{H}_p=&\sum_{\alpha=a,b}\int d\mathbf{r}[\frac{1}{2m_\alpha}|\nabla
\Phi_\alpha(\mathbf{r})|^2+V_a(\mathbf{r}) \rho_\alpha(\mathbf{r}) \\ &+
\frac{1}{2}c_0^\alpha\rho_\alpha^2(\mathbf{r})] +\int
d\mathbf{r}[c_0^{ab}\rho_a(\mathbf{r})\rho_b(\mathbf{r})] \end{array} $$ is the phase part,
$$ \begin{array}{rl} \mathcal{H}_s = & \sum_{\alpha=a,b}\frac{1}{2}\int
d\mathbf{r}[\frac{\rho_\alpha (\mathbf{r})| \nabla \mathbf{n}_\alpha(\mathbf{r})|^2}{m_\alpha}
+c_2^\alpha\mathbf{L}_\alpha^2(\mathbf{r})]\\ & +\int
d\mathbf{r}[c_2^{ab}\mathbf{L}_a(\mathbf{r}) \cdot\mathbf{L}_b(\mathbf{r})] \end{array} $$ is the spin part, and a spin-phase coupling part is negligible in the long-wavelength limit or when
$\mathbf{L}_\alpha =0$.
Hence the phase and spin degrees of freedom are decoupled and are described by the collective
variables $\{\rho_\alpha(\textbf{r}),\chi_\alpha(\textbf{r})\}$ and
$\{\textbf{n}_\alpha(\textbf{r}), \textbf{L}_\alpha(\textbf{r})\}$, respectively. Here $\mathcal{H}_p$ simply describes a
mixture of two scalar Bose gases, and is independent of the relative orientation of
$\textbf{n}_a$ and $\textbf{n}_b$. Henceforth we focus on the spin part $\mathcal{H}_s$.
First consider the uniform case $V_a=V_b=0$. In the ground state, $\rho_a$ and $\rho_b$ are
both constants and the total energy
\begin{eqnarray}
E =& {\cal V} (
\frac{1}{2}c_2^a
\mathbf{L}_a^2+\frac{1}{2}c_2^b\mathbf{L}_b^2+
c_2^{ab}\mathbf{L}_a\cdot\mathbf{L}_b \nonumber \\
& +
\frac{1}{2}c_0^a \rho_a^2+\frac{1}{2}c_0^b\rho_b^2+ c_0^{ab}\rho_a\rho_b), \end{eqnarray}
where ${\cal V}$ is the volume of the system. Minimization of $E$ implies that in the ground
state, $\Phi_a$, $\Phi_b$, $\textbf{n}_a$ and $\textbf{n}_b$ are
all position independent, while $\textbf{L}_a=\textbf{L}_b=0$, implying that the total spin is also $0$.
The mean-field ground state is $|N_a,\mathbf{n}_a\rangle \otimes
|N_b,\mathbf{n}_b\rangle$, where each species is in its own spin nematic state uncorrelated with the other species. Here $|N_\alpha,\mathbf{n}_\alpha\rangle$
denotes the state in which the spin director of each atom of species $\alpha$ is aligned along the direction of
$\mathbf{n}_\alpha$.
There are no constraints on $\textbf{n}_a$ and $\textbf{n}_b$, hence the ground state
manifold becomes $\frac{S^1\times S^2}{Z_2}\otimes \frac{S^1\times S^2}{Z_2}$, which possesses
a huge degeneracy.
\section{Quantum Fluctuations \label{qf}}
In the following we show that this degeneracy is lifted by zero-point quantum fluctuations. Denoted by $\textbf{n}_a^0$ and $\textbf{n}_b^0$, the spin directors of the two species in a mean-field symmetry-breaking ground state satisfy the relation $\textbf{n}_a^0 \cdot \textbf{n}_b^0 =
\cos\theta $. Let us arbitrarily set $\textbf{n}_a^0 =\textbf{e}_z$, $\textbf{n}_b^0 =\textbf{e}_{z'}$ and $\textbf{e}_z\cdot \textbf{e}_{z'}=\cos \theta$, as depicted in Fig~\ref{fig1}. The degeneracy implies that $\theta$ is arbitrary. The range of $\theta$ is
limited to $-\pi/2 \leq \theta < \pi/2$ because of $Z_2$ symmetry.
\begin{figure}
\begin{center}
\scalebox{0.6}{\includegraphics[-20pt,5pt][523pt,260pt]{gg.eps}}
\caption{ \label{fig1} The diagram on the left shows a typical mean-field configuration of the two nematic vectors $\textbf{n}_a^0=\textbf{e}_z$ and $\textbf{n}_b^0=\textbf{e}_{z'}$, with an arbitray angle $\theta$. The diagram on the right shows the small fluctuation $\mathbf{u}_z$ around $\textbf{n}_a^0$. The fluctuation $\mathbf{u}_b$ of $\textbf{n}_b$ in the $x'y'z'$ frame is similar. The coordinate systems are set such that $\textbf{e}_y=\textbf{e}_{y'}$. }
\end{center}
\end{figure}
Consider
\begin{equation}
\textbf{n}_\alpha = \textbf{n}_{\alpha}^0+\textbf{u}_\alpha,
\end{equation}
for $\alpha =a,b$, where quantum fluctuations are
$$ \textbf{u}_a \simeq\frac{\psi_{ax}}{\sqrt{\rho_a}}\textbf{e}_x+
\frac{\psi_{ay}}{\sqrt{\rho_a}}\textbf{e}_y,$$
$$\textbf{u}_b\simeq
\frac{\psi_{bx}}{\sqrt{\rho_b}}\textbf{e}_{x'}+
\frac{\psi_{by}}{\sqrt{\rho_b}}\textbf{e}_{y},$$ with unit vectors $\textbf{e}_x$, $\textbf{e}_y$, and
$\textbf{n}_a^0$ forming a Cartesian coordinate system with $\textbf{e}_{x'}$, $\textbf{e}_{y}$,
and $\textbf{n}_b^0 $ forming another one, as shown in Fig.~\ref{fig1}.
We find that
\begin{equation}
\mathcal{H}_s= \sum_{i=x,y} \mathcal{H}_{si},
\end{equation}
where
\begin{eqnarray}
\mathcal{H}_{si}=&\sum_{\alpha,\mathbf{k}}\frac{k^2}{2m_\alpha}
\psi_{\alpha i,\mathbf{k}}^{\dagger}\psi_{\alpha i,\mathbf{k} }+\frac{1}{2}c_2^\alpha\rho_\alpha\sum_{\alpha,\mathbf{k}}[2\psi_{\alpha i,\mathbf{k}}^{\dagger}\psi_{\alpha i,\mathbf{k} }\nonumber \\ &-(\psi_{\alpha i,\mathbf{k} }^{\dagger}\psi_{\alpha i, -\mathbf{k}}^{\dagger}+H.c.)]
+ c_2^{ab}\sqrt{\rho_a\rho_b} \zeta_i(\theta) \nonumber \\
&\sum_{\mathbf{k}}(\psi_{a i,\mathbf{k} }^{\dagger}\psi_{b i,\mathbf{k} }-\psi_{a i,\mathbf{k} }^{\dagger}\psi_{b i,-\mathbf{k} }^{\dagger}+H.c.),
\end{eqnarray}
written in terms of momentum $\mathbf{k}$, with $k\equiv |\mathbf{k}|$, $\zeta_x(\theta) =\cos\theta$, and $\zeta_y(\theta)=1$. Here
$\mathcal{H}_{si}$ depends only on $\psi_{\alpha i}$, $\mathcal{H}_{sx}$ depends on $\theta$, and $\mathcal{H}_{sy}$ is independent of $\theta$.
By performing a Bogoliubov transformation,
one obtains
\begin{equation}
\label{diag}
\mathcal{H}_{s i}=\sum_{\lambda=\pm,k} \omega_{i\lambda, k} (A_{i\lambda,k}^{\dag}A_{i\lambda,k}+\frac{1}{2}),
\end{equation}
where $A_{i\lambda,k}$ is some Bosonic operator and
\begin{eqnarray}
\omega_{i\pm,k}^2= & \frac{1}{2}
\big( \epsilon_{ak}^2+\epsilon_{bk}^2 \pm [(\epsilon_{ak}^2+\epsilon_{bk}^2)^2
-4E_{ak}E_{bk}(4g_ag_b \nonumber \\ & -4g_{ab}^2\zeta^2(\theta)+2g_aE_{bk}+2g_bE_{ak}+
E_{ak}E_{bk})]^{1/2} \big), \nonumber \\
\end{eqnarray}
with $E_{\alpha k} \equiv \frac{k^2}{2m_\alpha}$, $\epsilon_{\alpha k}^2
=E_{\alpha k}^2+2g_\alpha E_{\alpha k}$, $
g_\alpha \equiv c_2^\alpha\rho_\alpha$, and $g_{ab} \equiv c_2^{ab}\sqrt{\rho_a\rho_b}$.
Note that $\omega_{x\pm,k}$ depends on both $g_{ab}$ and $\theta$, while $\omega_{y\pm,k}$ depends on $g_{ab}$ but is independent of $\theta$.
\section{Fluctuation-induced Locking \label{lock}}
According to the spectra of ${\cal H}_{sx}$ and ${\cal H}_{sy}$ obtained above, we know that the quantum fluctuations lead to a $\theta$-dependent zero-point energy
\begin{equation}
{\cal E}_{0}(\theta)=E_0(\theta)+ E_0(\theta=0), \label{e00}
\end{equation}
where on the righthand side, the first term $E_0(\theta)$
is the zero-point energy of ${\cal H}_{sx}$, with
\begin{equation}
E_0(\theta)=\frac{1}{2} \sum_k \Omega_{k}(\theta),
\end{equation}
where
\begin{eqnarray}
\Omega_{k}(\theta) &\equiv&
\omega_{x+,k}+\omega_{x-,k} \nonumber \\
&= &
\big[\epsilon_{ak}^2+\epsilon_{bk}^2+2E_{ak}^{1/2}E_{bk}^{1/2}
(4g_ag_b-4g_{ab}^2\cos^2\theta \nonumber \\
&&+2g_a E_{bk}+2g_b E_{ak}
+E_{ak}E_{bk})^{1/2}\big]^{1/2},
\end{eqnarray}
which reaches its minimum at $\theta=0$. The second term $E_0(\theta=0)$ on the righthand side of (\ref{e00}), which is $\theta$-independent, is the zero-point energy of ${\cal H}_{sy}$.
Hence the total zero-point energy ${\cal E}_{0}(\theta)$ also reaches its minimum at $\theta=0$.
Therefore, in the ground state, $\textbf{n}_a^0$ and $\textbf{n}_b^0$ are actually locked in the low-energy limit, that is, they tend to align in the same direction. This is in contrast to what was suggested by the mean-field analysis.
For large $k$, the $\theta$-dependent part of $\Omega_{k}$ behaves as
$$-\frac{ g_{ab}^2 m_am_b}{k^2(m_a+m_b)}\cos^2\theta+O(\frac{g_{ab}^4}{k^6}), $$ thus the $\theta$-dependent part of the energy has ultraviolet divergence. This divergence originates from the use of contact interaction, which fails at short range or large momentum. It can be removed by introducing
a momentum cutoff or by the renormalization of interaction strengths, that is, by evaluating the ground-state energy in terms of the renormalized
quantities $c^{ab}_{0,r}$ and $c^{ab}_{2,r}$, which are directly related to the experimentally observed scattering lengths and correspond to the bare quantities $c^{ab}_0$ and $c^{ab}_2$ respectively.
It is known that $c_{0}^{ab}=\frac{2U_2+U_0}{3}$ and $c_{2}^{ab}=\frac{U_2-U_0}{3}$, where $U_0$ and $U_2$ are the interaction strengths
for the total spin $F=0$ and $2$ channels, respectively~\cite{luo}. Accordingly $c_{0,r}^{ab}=(2U_{2,r}+U_{0,r})/3$ and
$c_{2,r}^{ab}=(U_{2,r}-U_{0,r})/3$, with $$U_ {F,r}=\lim_{k\rightarrow 0} \langle \mathbf{k}',F |\hat{T}| \mathbf{k},F\rangle,$$ with $k=|\mathbf{k}|= |\mathbf{k}'|$, given by the zero-energy $\hat{T}$-matrix element for two-body scattering~\cite{ueda}. The
Lippman-Schwinger equation reads $$\hat{T}=\hat{U_F}+\hat{U_F}G_0\hat{T},$$ with $\hat{U_F}=U_F\delta(\mathbf{r})$ is the two-body potential and $$G_0(k)=-\frac{2M}{k^2},$$ where $$M \equiv \frac{m_am_b}{m_a+m_b}, $$ is the zero-energy Green's function for the relative motion of atoms $a$ and $b$. Consequently, $\frac{1}{U_{F,r}}=\frac{1}{U_{F}}+\int d^3k\frac{2M}{k^2}$. Since the divergence
occurs in the second order of interaction strengths, we expand the above formula to second order and obtain $$U_F= U_{F,r}+U^2_{F,r}\int d^3k\frac{2M }{k^2}.$$
Therefore, \begin{eqnarray}
c_0^{ab}&=&\frac{2U_2+U_0}{3}\nonumber \\ &=& \frac{2U_{2,r}+U_{0,r}}{3} +\frac{2U^2_{2,r}+U^2_{0,r}}{3}\int d^3k\frac{2M}{k^2} \nonumber \\
&=& c_{0,r}^{ab}+[(c_{0,r}^{ab})^2+2(c_{2,r}^{ab})^2]\int d^3k\frac{2M }{k^2}, \label{c0}
\end{eqnarray}
which is essential for the cancellation of the divergence.
Renormalization effect should be considered for all the terms of $c_0^{ab}$ and $c_2^{ab}$, including those in the mean field energy, in fluctuations of ${\cal H}_p$ and ${\cal H}_s$. In the mean field energy, by substituting (\ref{c0}) for $c_0^{ab}\rho_a\rho_b$ in the mean field energy, it can be seen that it becomes $c_{0,r}^{ab}\rho_a\rho_b$ after the $(c_{0,r}^{ab})^2$ term cancels the
divergent term in the Bogliubov ground state energy of $H_p$, while $(c_{2,r}^{ab})^2$ term cancels the divergent $(c_2^{ab})^2$ terms in ${\cal E}_0(\theta)$, with exact cancellation at $\theta =0$.
Therefore, the zero-point energy ${\cal E}_{0}(\theta)$ should be regularized by using
\begin{equation}
E_0(\theta)=\frac{{\cal V}}{2} \int \frac{d^3 k}{(2\pi)^3}[ \Omega_{k}(\theta) -\frac{\partial \Omega_{k}(\theta) }{\partial g_{ab}^2}g_{ab}^2 ],
\end{equation}
where the summation has been replaced with an integral.
It can be shown that after the
substraction, $\frac{\partial E_0}{\partial\theta}|_{\theta=0}=0$ and $\frac{\partial^2
E_0}{\partial\theta^2}|_{\theta=0}>0$ still hold, thus $E_0(\theta)$ remains minimal at
$\theta=0$.
Without a loss of the generality, we focus on the case $g_a= g_b= g$. By introducing
$k=k_0x$, where $k_0 \equiv \sqrt{2gM}$ is a characteristic momentum, we rewrite
$E_0(\theta) ={\cal V} k_0^3g I (\frac{g_{ab}}{g},\cos^2\theta)$, where
$$I (\frac{g_{ab}}{g},\cos^2\theta) \equiv \int \frac{d^3 x}{(2
\pi)^3}f\Big(\frac{g_{ab}}{g},\cos^2\theta,x\Big)$$ is dimensionless and $f$ is also dimensionless and depends on $g_{ab}/g$, $\cos^2\theta$, and the dimensionless quantity $x$. Hence
\begin{equation}
E_0(\theta)= g \sqrt{2M^3} N_\alpha \sqrt{\rho_\alpha (c_2^\alpha)^3} I
(\frac{g_{ab}}{g},\cos^2\theta),
\end{equation}
where $\sqrt{\rho_\alpha (c_2^\alpha)^3}$ is analogous to the Lee-Huang-Yang parameter in
dilute gas theory~\cite{lhy}.
The behavior of $E_0(\theta)$ for various values of $g_{ab}/g$ was numerically investigated and is shown in Fig.~\ref{fig2}.
The result indicates that $E_0(\theta)$
increases with $g_{ab}$ because of the enhancement of spin fluctuations.
The zero-point energy plays the role of an effective potential and can strongly
influence the coherent spin dynamics.
Although the mean-field spin dynamics has already been studied in quite a few
laboratories~\cite{Schmaljohann04,Chang05,Higbie05,Stamper12,Widera05}, the macroscopic quantum spin dynamics driven
by microscopic quantum fluctuations so far has not yet been explored in
experiments and remains to be observed. The phenomenon studied here provides motivation and a different venue in which to understand fluctuation-induced dynamics.
\begin{figure}
\begin{center}
\includegraphics[width=10cm]{s3.eps}
\caption{ \label{fig2} Zero-point energy $E_0(\theta)$ as a function of the angle $\theta$ between the spin directors $\mathbf{n}_a$ and $\mathbf{n}_b$. For simplicity, the dimensionless quantity $[E_0(\theta)-E_0(0)]/g$ is shown as the vertical coordinate. Note that $\theta$ is equivalent to $\theta+\pi$ because of the $Z_2$ symmetry. Here $g_{ab} \equiv
c_2^{ab}\sqrt{\rho_a\rho_b}$ and
$g_\alpha \equiv c_2^\alpha\rho_\alpha$, with $c_2^{\alpha}=\frac{4 \pi \hbar^2 \Delta
a_a}{m_a} $, $\Delta a_\alpha$ being the difference between the triplet and singlet
scattering lengths for atoms of species $\alpha$. In addition, $g_b=g_a=g$ is assumed without loss of generality. The parameter values are set as $\Delta a_a=0.1$nm, $\rho_a=10^{15}{\rm cm}^{-3}$,
$m_b/m_a=3.3$, and $N_a=10^4$. }
\end{center}
\end{figure}
\section{Fluctuation-induced spin dynamics \label{spind}}
In the present case, the effective Hamiltonian that controls the fluctuations of the spin
directors is
\begin{equation} \label{homo}
\mathcal{H}_{eff}=\sum_{\alpha }\frac{c_2^\alpha}{2\Omega}\mathbf{l}_{\alpha
}^2+\frac{c_2^{ab}}{\Omega }\mathbf{l}_{a}\cdot \mathbf{l}_{b}+E_0(\theta)+E_0(0).
\end{equation}
where $\textbf{l}_\alpha \equiv \Omega\textbf{L}_\alpha$. Defining the center-of-mass
quantities
$\mathbf{l}=\mathbf{l}_a+\mathbf{l}_b$ and
$$
\mathbf{n}=\frac{(c_2^b-c_2^{ab})\textbf{n}_{a}+(c_2^a-c_2^{ab})
\textbf{n}_{b}}{c_2^a+c_2^b-2c_2^{ab}},$$
and the relative quantities
$$\mathbf{l}_r=\frac{(c_2^a-c_2^{ab})\textbf{l}_{a}-(c_2^b-c_2^{ab})
\textbf{l}_{b}}{c_2^a+c_2^b-2c_2^{ab}}$$ and
$\mathbf{n}_r=\mathbf{n}_a-\mathbf{n}_b$, we can rewrite $\mathcal{H}_{eff}$ as
$\mathcal{H}_{eff} = \mathcal{H}_{c}+\mathcal{H}_{r}$, where
$$\mathcal{H}_{c}=\frac{1}{2\Omega}\frac{c_2^ac_2^b-(c_2^{ab})^2}{c_2^a+c_2^b-2c_2^{ab}}
\mathbf{l}^2+E_0(0)$$ is the center-of-mass part describing a free rotor and
$\mathcal{H}_{r}=\frac{1}{2\Omega}(c_2^a+c_2^b-2c_2^{ab})
\mathbf{l}_r^2+E_0(\theta)$ is the relative part. Note that
$\cos{\theta}=1-\textbf{n}_r^2/2$ and $\mathcal{H}_{c}$ and $\mathcal{H}_{r}$ are decoupled.
Let us focus on the relative motion and consider small oscillations around the minimum
$\theta=0$. To the lowest order, we can write $\textbf{n}_r=q_x\textbf{e}_x+q_y\textbf{e}_y$
and $\textbf{l}_r=l_{rx}\textbf{e}_x+l_{ry}\textbf{e}_y$, with $[q_ i,
l_{rj}]=i\epsilon_{ij}$ ($i,j=x,y$). Then up to a constant,
\begin{equation} \label{relaive}
\mathcal{H}_{r}= \sum_{i=x,y} [\frac{c_2^a+c_2^b-2c_2^{ab}}{2\Omega}
l_{ri}^2+\frac{K}{2}q_i^2].
\end{equation}
where $K=-\frac{\partial E_0}{\partial \cos\theta}|_{\theta=0}$. Hence $\mathcal{H}_{r}$ describes
two independent and identical harmonic oscillators, both with frequency
$$\omega_0\equiv \sqrt{\frac{(c_2^a+c_2^b-2c_2^{ab})K}{\Omega}}=
\sqrt{(\frac{g_a}{N_a}+\frac{g_b}{N_b}-2\frac{g_{ab}}{\sqrt{N_aN_b}})K}.$$
For typical values $g_a\sim g_b\sim 100$ Hz (in the unit of $\hbar$), $N_a\sim N_b\sim 10^4$,
and $g_{ab}/g=0.6$, $K$ can be numerically estimated as $\sim 6g_a$. The corresponding
frequency of the oscillation about the locked position is about $2$Hz. It can be substantially enhanced, even by a few orders of magnitude, when an optical
lattice is applied and the amplitude of fluctuation is tuned~\cite{song08}. More investigations are needed to address this circumstance.
The oscillation of the spin directors results in the oscillation of occupation numbers in the
Zeeman sublevels. As an example, let us consider
the case $\mathbf{l}=\mathbf{l}_a+\mathbf{l}_b=0$ so that $ \mathbf{n}$ is independent of time.
Then $$\textbf{n}_{a}=\mathbf{n}+
\frac{c_2^a-c_2^{ab}}{c_2^a+c_2^b-2c_2^{ab}}\textbf{n}_r$$
and $$\textbf{n}_{b}=\mathbf{n}-
\frac{c_2^b-c_2^{ab}}{c_2^a+c_2^b-2c_2^{ab}}\textbf{n}_r. $$ Thus the spin states of
species $a$ and $b$ in the Zeeman basis state of $m=\pm 1$ are $$\xi_{a \pm 1}=
\frac{c_2^a-c_2^{ab}}{\sqrt{2}(c_2^a+c_2^b-2c_2^{ab})}(iq_y\mp q_x)$$ and $$\xi_{b \pm 1}=-
\frac{c_2^b-c_2^{ab}}{\sqrt{2}(c_2^a+c_2^b-2c_2^{ab})}(iq_y\mp q_x), $$ while the spin states
of species $a$ and $b$ in the Zeeman basis state of $m=0$ are both $\mathbf{n}$. Since $q_x$
and $q_y$ both oscillate with frequency $\omega_0$, the
occupation number $N_\alpha |\xi_{\alpha \pm 1}|^2$ oscillates with frequency $2\omega_0$. The occupation numbers may be probed by using, say, an optical cavity~\cite{cui}.
\section{Locking in a trap \label{trap}}
Now we examine how the locking effect survives a trapping potential, which is an
experimental necessity. Supposing that the potential has the harmonic form and the clouds of the two species have the
same size $R$, then we have
\begin{equation}\label{dense}
\rho_\alpha=A_\alpha (R^2-r^2),
\end{equation}
where $A_\alpha$ is a positive constant.
The spin dynamics is determined by the Heisenberg equations $$i\partial_t
\textbf{n}_\alpha(\textbf{r})=[\textbf{n}_\alpha(\textbf{r}), \mathcal{H}_s], $$ $$i\partial_t
\textbf{L}_\alpha(\textbf{r})=[\textbf{L}_\alpha(\textbf{r}), \mathcal{H}_s].$$ For small
fluctuations, we can impose the commutation relations $$
[\mathbf{L}^i_\alpha(\mathbf{r}), \mathbf{n}^j_\beta(\mathbf{r}')]=
i\delta_{\alpha\beta}\epsilon^{ijk}\mathbf{n}^k_\alpha(\mathbf{r})
\delta(\mathbf{r}-\mathbf{r}'), $$
$$[\mathbf{L}^i_\alpha(\mathbf{r}), \mathbf{L}^j_\beta(\mathbf{r}')]
= i\delta_{\alpha\beta}\epsilon^{ijk}\mathbf{L}^k_\alpha(\mathbf{r})
\delta(\mathbf{r}-\mathbf{r}'),$$
where $i, j, k=x, y, z$. One obtains
\begin{equation}
\begin{array}{rcl}
{\partial}_t\mathbf{n}_\alpha & = & c_2^\alpha(\mathbf{n}_\alpha \times\mathbf{L}_\alpha)
+c_2^{ab}(\mathbf{n}_\alpha\times\mathbf{L}_{\beta}), \\
{{\partial}}_t\mathbf{L}_\alpha & = & \frac{1}{m_\alpha}
\mathbf{n}_\alpha\times\nabla \cdot (\rho_\alpha\nabla\mathbf{n}_\alpha) +
c_2^{ab} \mathbf{L}_\beta\times\mathbf{L}_\alpha,
\end{array} \label{spin}
\end{equation}
where $\alpha\neq\beta$ and $\nabla^2 \sqrt{\rho}$ terms are neglected for large clouds. In
the Thomas-Fermi approximation, the spin structure for the ground state remains unaffected, thus
the fluctuating directors can be generally written as $$\textbf{n}_a= u_{ax}
\textbf{e}_x+u_{ay}\textbf{e}_y+ \sqrt{1-u_{ax}^2-u_{ay}^2} \textbf{n}_a^0$$ and
$$\textbf{n}_b=u_{bx} \textbf{e}_x'+
u_{by'}\textbf{e}_y+\sqrt{1-u_{bx}^2-u_{by}^2}\textbf{n}_b^0.$$ To first order of $u_{\alpha
i}$ and $L_{\alpha i}$,
\begin{equation}\label{en}
\begin{array}{rcl}
\partial_t u_{\alpha y}&=&c_2^\alpha L_{ \alpha x}+c_2^{ab}L_{\beta x}\cos \theta, \\
\partial_t L_{\alpha x}&=&\frac{1}{m_\alpha}\nabla\cdot(\rho_\alpha \nabla u_{\alpha y}). \\
\end{array}
\end{equation}
The two eigenfrequencies are given by
\begin{equation}
\omega_{\pm}^2=\frac{f_{nl}}{2}\big[
\nu_a^2+\nu_b^2\pm[(\nu_a^2-\nu_b^2)^2
+4\nu_{ab}^2\nu_{ba}^2\cos^2\theta]^{1/2}
\big],\end{equation}
where $f_{nl}\equiv 2n^2+2nl+3n+l$, $l$ is the angular quantum number, $2n$ is the order of a polynomial of even powers describing the radial wave function~\cite{stringari}, $\nu_a^2=\frac{Ac_2^a}{m_a}$,
$\nu_b^2=\frac{Bc_2^b}{m_b}$, $\nu_{ab}^2=\frac{Bc_2^{ab}}{m_b}$, and
$\nu_{ba}^2=\frac{Ac_2^{ab}}{m_a}$.
Therefore
$\omega_{+}+\omega_{-}=(f_{nl})^{1/2}
[\nu_a^2+\nu_b^2+2(\nu_a^2\nu_b^2-\nu_{ab}^2\nu_{ba}^2\cos^2\theta
)^{1/2}]^{1/2}$,
which reaches its minimum at $\theta=0$. Hence the locking also occurs in a trap.
Now let us address the effect of an external magnetic field along the $z$ direction, the
presence of which breaks the full rotational symmetry down to $S^1$ symmetry, leading to a
nonzero mean-field value of $L_{\alpha z}$. It can be shown that in the mean-field
ground state~\cite{stenger}, each species $\alpha$ undergoes Bose-Einstein condensation with
the Zeeman basis wave function $(\psi_{\alpha,1}, 0, \psi_{\alpha,-1})^T$, hence the ground
state only possesses spin rotation symmetry $S^1\times S^1$, as the two spins can rotate
around the $z$ axis independently without changing the mean-field energy. Due to the Zeeman barrier,
the low energy dynamics is dominated by the phase fluctuations of $\psi_{\alpha,1}$ and
$\psi_{\alpha,-1}$, while $\psi_{\alpha,0}$ remains zero, thus the effective Hamiltonian
describes a mixture of two pseudospin-$\frac{1}{2}$ Bose gases~\cite{shi1}. For low enough magnetic fields, the
spin fluctuation can overcome the Zeeman barrier and restore $S^2$ symmetry~\cite{zhou1}, consequently the locking persists.
\section{Summary \label{summ}}
Note that the spin directors are macroscopic collective variables of the spinor Bose gases, due
to Bose-Einstein condensation. We have shown that in our system, microscopic quantum fluctuations dramatically change the nature of spontaneous symmetry breaking.
Two interacting macroscopic systems undergo spontaneous symmetry breaking in a correlated way, and consequently the two macroscopic collective variables are locked.
The symmetry-breaking states of the system are
$|N_a,\mathbf{n}\rangle \otimes |N_b,\mathbf{n}\rangle$,
where the spin directors of the two species are locked to be ${\bf n}$ along an arbitrary direction.
They are in contrast to states $|N_a,\mathbf{n}_a\rangle \otimes |N_b,\mathbf{n}_b\rangle$, as suggested by the simple mean-field analysis, where $\mathbf{n}_a$ and $\mathbf{n}_b$ are arbitrary and independent of each other.
To summarize, by considering a mixture of two distinct species of spin-1 atoms with
interspecies spin exchange, we have shown that the zero-point quantum fluctuations lift the ground
state degeneracy suggested by the mean field theory and lead to the locking between the spin directors of the two species under the experimentally realistic conditions.
This is a type of quantum phenomenon in which the microscopic quantum fluctuations fundamentally control the macroscopic collective phenomenon, by changing the very nature of symmetry breaking.
\acknowledgments
This work was supported by the National Science Foundation of China (Grant No. 11374060) and
NSERC (Canada). F.Z. was also supported by Canadian Institute for Advanced Research.
|
\section{Introduction}
Considering the richness of Ramsey theory and the great interest in random graphs, it is natural to consider Ramsey properties of random graphs.
The study of random Ramsey theory has proved particularly useful in the establishment of upper bounds on the size Ramsey number.
For graphs $G,F,H$, we write $G\rightarrow (F,H)$ if for every red-blue colouring of the edges of $G$, there is either a red $F$ or a blue $H$. If $F,H$ are isomorphic, we use instead the notation $G\rightarrow H$. The size Ramsey number, denoted by $\hat{r}(H)$ is defined to be $\hat{r}(H)=\min\{|E(G)|: G\rightarrow H\}$.
In \cite{beck}, disproving a conjecture of Erd\H{o}s \cite{erdos}, Beck showed that $\hat{r}(P_n)\le 900 n$. In \citep{bela} Bollob\'as noted a slightly better bound, and recently Dudek and Pra\l{}at \citep{pralat} gave an elementary proof of the bound $\hat{r}(P_n)\le 137n$. They actually proved that w.h.p.,~$G(n,\alpha/n)\rightarrow P_{\beta n}$ for some constants $\alpha, \beta$.
They raised the question of determining the maximum $l$ such that $G(n,p)\rightarrow P_l$, where $pn\rightarrow \infty$.
Inspired by the well known result of Gerencs\'er and Gy\'arf\'as \cite{gyarfas} which says that $K_n\rightarrow P_{2n/3}$, they ask if $G(n,p)\rightarrow P_l$ for some $l=n(2/3+o(1))$.
Our main result answers this question in the affirmative.
\begin{thm}\label{thm_random_path_ramsey}
Let $0<p=p(n)<1$ and assume that $pn\rightarrow\infty$.
Then w.h.p., $G(n,p)\rightarrow P_l$ for some $l=(2/3+o(1))n$.
\end{thm}
This result is essentially best possible since there is a $2$-colouring of the edges of $K_n$ such that the longest monochromatic path is of length $\lceil 2n/3 +1\rceil$. To see this, divide the vertex set of $K_n$ into two sets $A,B$ such that $|A|=\lfloor n/3 \rfloor$, let the edges spanned by $B$ be coloured red and colour the other edges blue.
In order to prove Theorem \ref{thm_random_path_ramsey}, we prove the following extension of the result in \cite{gyarfas} to graphs with a large number of edges.
\begin{thm}\label{thm_density_path_ramsey}
Let $0\le\varepsilon\le 1/4$, $k\ge l$ and let $G$ be a graph on $n\ge k+\lfloor(l+1)/2\rfloor+150\sqrt{\varepsilon} k$ vertices, with at least $(1-\varepsilon)\binom{n}{2}$ edges.
Then $G\rightarrow (P_{k+1},P_{l+1})$.
In particular, given $0\le \varepsilon<1/90000$, for every graph $G$ on $n$ vertices and at least $(1-\varepsilon)\binom{n}{2}$ edges, $G\rightarrow P_k$ where $k=(2/3-100\sqrt{\varepsilon})n$.
\end{thm}
Theorem \ref{thm_density_path_ramsey} is a consequence of the following similar result, in which we consider graphs with large minimum degree rather than high density.
\begin{thm}\label{thm_min_deg_path_ramsey}
Let $0<\varepsilon<1/4$, $k\ge l$ and let $G$ be a graph on $n\ge k+\lfloor(l+1)/2\rfloor+60\varepsilon k$ vertices with minimum degree at least $(1-\varepsilon)n$. Then $G\rightarrow (P_{k+1},P_{l+1})$.
\end{thm}
Note that it is easy to deduce Theorem \ref{thm_density_path_ramsey} from Theorem \ref{thm_min_deg_path_ramsey}.
By an averaging argument, it suffices to prove the assertion for $n=k+l/2+150\sqrt{\varepsilon}$.
By removing at most $\sqrt{\varepsilon}n$ vertices, we obtain a graph on $n'\ge (1-\sqrt{\varepsilon})n$ vertices and minimum degree at least $(1-\sqrt{\varepsilon})n\ge (1-\sqrt{\varepsilon})n'$ vertices.
One can check that $(1-\sqrt{\varepsilon})n\ge k+l/2+100\sqrt{\varepsilon}k$, so the assertion of Theorem \ref{thm_density_path_ramsey} follows from Theorem \ref{thm_min_deg_path_ramsey}.
For the second part, it is easy to check that when $k=l=(2/3-100\sqrt{\varepsilon})n$ and $150\sqrt{\varepsilon}\le 1/2$, it follows that $n\ge k+l/2+150\sqrt{\varepsilon}$.
The rest of the paper is organised as follows.
In Section \ref{sec_dense} we prove Theorem \ref{thm_min_deg_path_ramsey}. In
order to prove Theorem \ref{thm_random_path_ramsey}, we use the so-called sparse regularity lemma, due to Kohayakawa \citep{kohayakawa} and R\"odl (see \cite{conlon}). In Section \ref{sec_regularity} we state this result as well as some necessary notation.
We prove Theorem \ref{thm_random_path_ramsey} in Section \ref{sec_random} and finish with some concluding remarks in Section \ref{sec_conclusion}.
Throughout the paper we omit floor and ceiling signs whenever they do not affect the arguments.
\section{Path Ramsey number for dense graphs}\label{sec_dense}
In the proof of Theorems \ref{thm_min_deg_path_ramsey} and \ref{thm_random_path_ramsey} we use the following observation, from \cite{pralat} and \citep{Pokrovskiy}. For the sake of completeness, we prove it here.
\begin{lem}\label{lem_partition}
For every graph $G$ there exist two disjoint sets of vertices $U,W$ of equal sizes, such that there are no edges between them and $G\backslash (U\cup W)$ has a Hamiltonian path.
\end{lem}
\begin{proof}
In order to find sets with the desired properties, we apply the following algorithm, maintaining a partition of $V(G)$ into sets $U,W$ and a path $P$.
Start with $U=V(G), W=\emptyset$ and $P$ an empty path.
At every stage in the algorithm, do the following.
If $|U|\le|W|$, stop.
Otherwise, if $P$ is empty, move a vertex from $U$ into $W$ (note that $U\neq \emptyset$).
If $P$ is non-empty, let $v$ be its endpoint. If $v$ has a neighbour $u$ in $U$, put $u$ in $P$, otherwise move $v$ to $W$.
Note that at any given point in the algorithm there are no edges between $U$ and $W$.
Furthermore, the value $|U|-|W|$ is positive at the beginning of the algorithm and decreases by one at every stage, thus at some point the algorithm will stop and produce sets $U,W$ with the required properties.
\end{proof}
Occasionally it is easier to use the following immediate consequence of Lemma \ref{lem_partition}.
\begin{cor}\label{cor_partition}
Let $G$ be a balanced bipartite graph on $n$ vertices with bipartition $V_1,V_2$, which has no path of length $k$. Then there exist $X_i\subseteq V_i$ such that $|X_1|=|X_2|\ge (n-k)/4$ and $G$ has no edges between $X_1$ and $X_2$.
\end{cor}
\begin{proof}
Let $U,W$ be as in Lemma \ref{lem_partition} and let $P$ be a Hamiltonian path in $G\backslash(U\cup W)$.
Note that $P$ must alternate between $V_1$ and $V_2$, thus $|V(P)\cap V_1|=|V(P)\cap V_2|$ (it follows from the assumptions that the number of vertices in $P$ is even).
Denote $U_i=U\cap V_i$, $W_i=W\cap V_i$, for $i=1,2$ and assume that $|U_1|\ge |U_2|$.
It follows that $|U_1|+|W_1|=|U_2|+|W_2|$.
Thus, using the fact that $|U|=|W|$, we have that $|U_1|=|W_2|\ge |U|/2\ge (n-k)/4$.
Set $X_1=U_1$ and $X_2=W_2$.
\end{proof}
We now prove Theorem \ref{thm_min_deg_path_ramsey}.
\begin{proof} [Proof of Theorem \ref{thm_min_deg_path_ramsey}]
We prove the theorem by induction on $k$.
Clearly, if $k=1$ the claim holds.
Given $k>1$, let $G$ be a graph on $n\ge k+\lfloor (l+1)/2\rfloor+100\varepsilon k$ vertices, with minimum degree at least $(1-\varepsilon)n$, and consider a red-blue colouring of the edges of $G$.
If $k>l$ then by induction there is either a red $P_k$ or a blue $P_{l+1}$; in the latter case we are done.
If $k=l$ then by induction there is either a red or blue $P_k$.
Thus, without loss of generality there is a red path of length $k-1$, which we denote by $P=(v_1,\ldots,v_k)$. Let $U=V(G)\backslash V(P)$.
We note first that the assertion of Theorem \ref{thm_min_deg_path_ramsey} holds when $k\ge n(1/2-\varepsilon)$.
If there is no red $P_{k+1}$, then by Lemma \ref{lem_partition}, we can find disjoint sets $U,W$, of size at least $(n-k)/2$ such that there is no red edge between them.
Since $G$ has minimum degree at least $n(1-\varepsilon)$, we can greedily find a blue path of length at least $|U|+|W|-2\varepsilon n\ge n(1-2\varepsilon)-k\ge k$.
Thus we can assume that $k\ge 4n$, so every vertex in $G$ has at most $4\varepsilon k$ non neighbours.
Put $\delta=4\varepsilon$.
Note that we can assume that $\delta k\ge 1$, otherwise $G$ is a complete graph and Theorem \ref{thm_min_deg_path_ramsey} follows directly from \cite{gyarfas}.
We consider three cases.
\subsection{Case 1. $G[U]$ contains a blue path $Q$ of length $13\delta k$}
Let $Q_1$ be a maximal path extending $Q$ by alternating between vertices of $P$ and $U$ and which has both ends in $U$.
Let $U'=U\backslash V(Q_1)$ and $V'=V(P)\backslash V(Q_1)$.
Let $Q_2$ be a maximal path alternating between $U'$ and $V'$ which has both ends in $U$.
Denote the ends of $Q_i$ by $x_i,y_i$, for $i=1,2$.
We show that $|Q_1|+|Q_2|\ge l+3\delta k$.
Suppose this is not the case. In particular, $Q_1,Q_2$ do not cover $U$, so we can pick a vertex $z\in U\backslash (V(Q_1)\cup V(Q_2) )$.
Note that all but at most $3\delta k$ vertices of $P$ are adjacent to all of $x_1,x_2,z$. By our assumption on the lengths of $Q_1$ and $Q_2$, the number of vertices of $P$ which are included in one of $Q_1$ and $Q_2$ is at most $k/2 -5\delta k$, hence there exist vertices $v_i,v_{i+1}$ which are adjacent to all of $x_1,y_1,z$.
We assume that $v_i$ and $v_{i+1}$ have no common red neighbour in $x_1,x_2,z$ because otherwise we obtain a red $P_{k+1}$.
It follows that without loss of generality, $v_i$ is joined in blue to two of $x_1,y_1,z$, contradicting the maximality of $Q_1$ and $Q_2$.
Let $Q_2'$ be a subpath of $Q_2$ with ends $x_2',y_2'\in U$ satisfying $|Q_2|+|Q_1|=l+3\delta k$.
A similar argument to the above shows that without loss of generality there exist $v_i,v_{i+1}$ such that $x_1,y_1$ are blue neighbours of $v_i$ and $x_2',y_2'$ are blue neighbours of $v_{i+1}$.
Denote by $C_1$ and $C_2$ the blue cycles obtained by adding $v_i$ to $Q_1$ and $v_{i+1}$ to $Q_2$, and let $U_i=V(C_i)\cap U$.
Note that if there is any blue edge between $C_1$ and $C_2$ we obtain a blue path of length $l$, so we assume that no such edges exist. We can also assume that $|U_1|,|U_2|\ge 3\delta k$, otherwise one of $Q_1,Q_2$ has length at least $l$.
The number of vertices in $V(P)\backslash (V(C_1)\cup V(C_2))$ is at least $k/2+5\delta k$, hence there exists $j$ such that $v_j,v_{j+1}\notin V(C_1)\cup V(C_2)$.
If one of $v_j$ and $v_{j+1}$ has blue neighbours in both $U_1$ and $U_2$, we obtain a blue path of length $l$, so we can assume this is not the case.
Also, we assume that $v_j,v_{j+1}$ have no red common neighbour in either $U_1$ or $U_2$, because otherwise we obtain a red path of length $k$.
Thus, recalling that $v_j,v_{j+1}$ have at most $\delta k$ non neighbours in $G$, without loss of generality, $v_j$ is joined in red to all but $\delta k$ vertices of $U_1$, and $v_{j+1}$ is joined in red to all but $\delta k$ vertices in $U_2$.
Let $w_1\in U_1$ be any red neighbour of $v_j$. Since it is connected to all but at most $\delta k$ vertices of $U_2$ and these edges must all be red, $U_2$ contains a vertex $w_2$ which is a red neighbour of both $w_1$ and $v_{j+1}$.
We obtain a red path $v_1,\ldots,v_j,w_1,w_2,v_{j+1},\ldots,v_k$ of length $k$.
This finishes the proof of Theorem \ref{thm_density_path_ramsey} in the first case.
\subsection{Case 2. $l\le (1-13\delta) k$}
Let $Q_1$ be a maximal blue path alternating between $U$ and $P$ and having both ends in $U$ and similarly let $Q_2$ be a maximal blue path alternating between $U\backslash V(Q_1)$ and $V(P)\backslash V(Q_1)$.
As in the previous case, it can be shown that $|Q_1|+|Q_2|\ge l+3\delta k$.
Let $Q_2'$ be a subpath of $Q_2$ such that $|Q_1|+|Q_2'|=l+3\delta k$.
As before, there exists $j$ such that both $v_j,v_{j+1}\in V(P)\backslash (V(Q_1)\cup V(Q_2'))$ and they are joined in $G$ to all ends of the two paths. Thus the vertices $v_j,v_{j+1}$ can be used to extend $Q_1,Q_2'$ into blue vertex disjoint cycles $C_1,C_2$, whose sum of length is $l+3\delta k$ and each of which has length at least $3\delta k$.
The proof of Theorem \ref{thm_min_deg_path_ramsey} can now be finished as in the first case.
\subsection{Case 3. $l\ge (1-13\delta)k$ and $G[U]$ contains no blue path of length at least $13\delta k$}
We conclude from Lemma \ref{lem_partition} that there exist two disjoint sets $W_1,W_2\subseteq U$ of size $|W_1|=|W_2|\ge (1/2+3\delta) k/2$ with no blue edges between them.
Since every vertex in $G$ is adjacent to all but at most $\delta k$ vertices, we can greedily find a red path $Q$ in $U$ of length at least $|W_1|+|W_2|-2\delta k=(1/2 +\delta )k$.
Let $X$ be the set of the first and last $ (1/4+\delta/2)k$ vertices of $P$.
We assume that there is no red edge between $X$ and $Q$, because otherwise there is a red path of length $k$.
We can now greedily construct a blue path alternating between $X$ and $V(Q)$ of length at least $|X|+|Q|-2\delta k\ge k\ge l$.
\end{proof}
\section{Sparse regularity lemma}\label{sec_regularity}
We shall make use of a variant of Szemer\'edi's regularity lemma \cite{regularity} for sparse graphs, often referred to as the sparse regularity lemma, which was proved independently by Kohayakawa \cite{kohayakawa} and R\"odl (see \cite{conlon}).
Before stating the theorem, we introduce some notation.
Given two disjoint sets of vertices $U,V$ in a graph, we define the density $d_p(U,V)$ of edges between $U$ and $V$ with respect to $p$ to be
\begin{equation}
d_p(U,V)=\frac{e(U,V)}{p|U||V|},
\end{equation}
where $e(U,V)$ is the number of edges between $U$ and $V$.
We say that a bipartite graph with bipartition $U,V$ is $(\varepsilon, p)$-regular if for every $U'\subseteq U, V'\subseteq V$ with $|U'|\ge \varepsilon |U|, |V'|\ge \varepsilon|V| $ the density $d_p(U',V')$ satisfies $|d_p(U',V')-d_p(U,V)|\le \varepsilon $.
Given a graph $G$, a partition $V_1,\ldots,V_t$ of $V(G)$ is called an $(\varepsilon,p)$-regular partition if it is an equipartition (i.e. the sizes of the sets differ by at most one), and if all but at most $\varepsilon$ of the pairs $V_i,V_j$ induce an $(\varepsilon,p)$-regular graph.
Given $0<\eta,p<1, D\ge 1$, a graph $G$ is called $(\eta, p, D)$-upper-uniform if for all disjoints sets of vertices $U_1,U_2$ of size at least $\eta |V(G)|$, the density $d_p(U_1,U_2)$ is at most $D$. Note that random graphs are w.h.p.~upper uniform (with suitable parameters).
We are now ready to state the sparse regularity lemma of Kohayakawa and R\"odl.
\begin{thm}\label{thm_sparse_regularity}
For every $\varepsilon>0$, $t$ and $D>1$ there exist $\eta>0$ and $T$ such that for every $0\le p\le 1$, every $(\eta,p,D)$-upper-uniform graph admits an $(\varepsilon,p)$-regular partition into $s$ parts where $t\le s\le T$.
\end{thm}
We shall use a slightly stronger variant of \ref{thm_sparse_regularity}, namely the coloured version of the sparse regularity lemma.
\begin{thm}\label{thm_coloured_sparse_regularity}
For every $\varepsilon>0$, $t,l$ and $D>1$ there exist $\eta>0$ and $T$ such that for every $0\le p\le 1$, if $G_1,\ldots,G_l$ are $(\eta,p,D)$-upper-uniform graphs on vertex set $V$, there is an equipartition of $V$ into $s$ parts, where $t\le s\le T$, for which all but at most $\varepsilon$ of the pairs induce a regular pair in each $G_i$.
\end{thm}
\section{Path Ramsey number for random graphs}\label{sec_random}
Before turning to the proof of Theorem \ref{thm_random_path_ramsey}, we remark that a weaker result can be proved using elementary tools.
\begin{lem}
Let $0<p=p(n)<1$ and assume that $pn\rightarrow\infty$.
Then w.h.p., $G(n,p)\rightarrow P_l$ for some $l=(1/2+o(1))n$.
\end{lem}
\begin{proof}
Given $\alpha>0$, suppose $G$ can be coloured such that there is no monochromatic path of length $n(1/2-\alpha)$.
By Lemma \ref{lem_partition}, there exist disjoint sets $U,W$, both of size at least $n(1/2+\alpha)/2$ with no red edges between them.
Considering the graph $G[U,W]$, it follows from the same lemma that there exist disjoint sets $X\subseteq U,Y\subseteq W$ of size at least $\alpha n/2$, such that there are no blue edges between them. We conclude that there are no edges of $G$ between $X$ and $Y$.
But w.h.p.,~every two disjoint sets of at least $\alpha n/2$ vertices in $G$ have an edge between them.
This shows that w.h.p., in every $2$-colouring of $G$ there is a monochromatic path of length at least $n(1/2-\alpha)$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm_random_path_ramsey}]
Let $0<p<1$ be such that $pn\rightarrow \infty$ and let $\alpha>0$.
We show that w.h.p., for every $2$-edge-colouring of $G=G(n,p)$ there is a monochromatic path of length at least $(2/3-\alpha)n$.
We pick $\varepsilon>0$ small and $t$ large (taking $t=1/\varepsilon$ and $\varepsilon$ small enough such that $60\sqrt{\varepsilon}\le \alpha$ would do).
Let $\eta, T$ be the constants arising from the application of Theorem \ref{thm_coloured_sparse_regularity} with $\varepsilon,t,l=2,D=2$.
Note that $G$ is w.h.p.~$(\eta,p,2)$-upper-uniform. Thus, by Theorem \ref{thm_coloured_sparse_regularity}, given a $2$-edge-colouring of $G$, there exists an $(\varepsilon,p)$-regular partition $V_1,\ldots,V_s$ with $t\le s\le T$.
Again, w.h.p.,~the density of edges $d_p(V_i,V_j)$ is at least $1/2$.
Let $H$ be the auxiliary graph with vertex set $[s]$ where $ij$ is an edge iff $V_i,V_j$ induce a regular bipartite graph in both red and blue.
We colour an edge $ij$ in $H$ red if the red density $d_p(V_i,V_j)$ is at least $1/4$ and blue otherwise, so if $ij$ is blue then the blue density is at least $1/4$.
Since the partition $V_1,\ldots,V_s$ is $(\varepsilon,p)$-regular, the number of edges in $H$ is at least $(1-\varepsilon)\binom{s}{2}$.
It follows from Theorem \ref{thm_density_path_ramsey} that $H$ contains a monochromatic path $P$ on at least $l=(2/3-\delta)s$ vertices, where $\delta=50\sqrt{\varepsilon}$ (assuming $\varepsilon>0$ is small enough).
Denote by $i_1,\ldots,i_l$ the vertices of $P$.
Assuming without loss of generality that $P$ is red, we show that $G$ contains a red path of length at least $(2/3-\alpha)n$.
We divide each set $V_{i_j}$ into two sets $U_j,W_j$ of equal sizes, so $|U_j|=n/2s$.
Let $P_j$ be a longest red path in the bipartite graph $G[U_j,W_{j+1}]$. In the following claim we show that $P_j$ covers most vertices in $U_j\cup W_{j+1}$. We shall then show that consecutive paths $P_j,P_{j+1}$ can be connected without losing too many vertices, thus obtaining a red path in $G$ of the required length.
\begin{claim}\label{claim_path_regular_pair}
For every $1\le j\le l$, $P_j$ covers at least $1-4\varepsilon$ of the vertices of $U_j\cup W_{j+1}$.
\end{claim}
\begin{proof}
Suppose that for some $j$, $P_j$ covers at most $1-4\varepsilon$ of the vertices of $U_j\cup W_{j+1}$.
Set $U=U_j$ and $W=W_{j+1}$.
By Corollary \ref{cor_partition}, there exist sets $X\subseteq U, Y\subseteq W$ with $|X|=|Y|\ge \varepsilon|U|$, such that there are no red edges between $X$ and $Y$.
But by the regularity of the partition $V_1,\ldots,V_s$, the density $d_p(U,V)$ is within $\varepsilon $ of the density of red edges between $U$ and $W$, which is at least $1/4$. In particular, $G$ has a red edge between $X$ and $Y$, contradicting our assumption, so Claim \ref{claim_path_regular_pair} holds.
\end{proof}
We now show that the paths $P_1,\ldots, P_{l-1}$ can be joined to a path $Q$ without losing many of the vertices.
Let $X_j$ be the set of first $2\varepsilon |V_1|$ vertices of $P_j$ and similarly let $Y_j$ be the set of last $2\varepsilon |V_1|$ vertices of $P_j$.
Since the paths $P_j$ alternate between the sets $U_j,W_{j+1}$, we have that $|Y_j\cap V_{i_j}|,|X_{j+1}\cap V_{i_{j+1}}|\ge \varepsilon |V_1|$. It follows from the fact that $i_ji_{j+1}$ is a blue edge in $H$ that there is a blue edge between $Y_j$ and $X_{j+1}$.
Hence $G$ has a blue path $Q$ which contains all vertices of $V(P_1)\cup\ldots\cup V(P_{l-1})$ but at most $4\varepsilon |V_1|(l-1)$.
Using Claim \ref{claim_path_regular_pair}, we have that
$|P_j|\ge (1-4\varepsilon)|V_1|$, so
\begin{align*}
|Q|\ge &(1-8\varepsilon)(l-1)|V_1|=
(1-8\varepsilon)(s(2/3-\delta)-1)n/s\ge\\
& n(2/3-(\delta+1/t+6\varepsilon))\ge n(2/3-\alpha).
\end{align*}
This completes the proof of Theorem \ref{thm_random_path_ramsey}.
\end{proof}
\section{Concluding Remarks}\label{sec_conclusion}
It is easy to construct examples of graphs $G$ on $n$ vertices with $n\ge k+\lfloor(l+1)/2\rfloor+c\varepsilon k$ and at least $(1-\varepsilon)\binom{n}{2}$ edges which admits a red-blue colouring with no red $P_{k+1}$ or blue $P_{l+1}$ (e.g. by letting as many vertices as possible be isolated).
It may be interesting to determine the correct dependence of $n$ on $\varepsilon$ in Theorem \ref{thm_density_path_ramsey}. In particular, can $\sqrt{\varepsilon}$ be replaced by a factor of $\varepsilon$?
Similarly, it would be interesting to determine if the error term $c\varepsilon k$ in Theorem \ref{thm_min_deg_path_ramsey} can be replaced by $o(\varepsilon k)$.
|
\section{Introduction}
Conformal mapping techniques are powerful tools for analyzing two dimensional models, and have been applied to a variety of problems \cite{schinzinger2012conformal} including aerodynamics \cite{milne1996}, fluid flows \cite{milne1996,Batchelor2000}, free boundary problems \cite{Jeong1992,Jeong2007,howison1992} potential theory/electrostatics \cite{tikhonov1990}, elasticity \cite{sokolnikoff1956} and even computer vision \cite{sharon20062d}. Abstractly, a problem of interest in the ``physical plane" is transformed into a simpler problem in the ``pre-image plane" by an appropriately chosen complex analytic function, i.e. a conformal map \cite{nehari-conformal} relating the pre-image plane to the physical plane. In addition to generating analytic solutions, conformal mapping methods have also been applied in conjunction with numerical techniques to solve various problems of physical interest \cite{henrici1993}.
An important issue in applying numerical (discretized) conformal maps to two dimensional problems where the domain has sharp regions/near singularities is the phenomenon of {\em crowding}. Crowding refers to the tendency for image nodes to accumulate in the physical plane near regions of high boundary curvature and leave low curvature regions extremely coarsely represented. Crowding plagues all standard numerical conformal mapping methods when used to generate image domains with highly curved boundaries \cite{Porter2005,Wegmann2005},
and there are specialized algorithms to mitigate the effects of crowding for conformal mappings to {\em known} polygonal domains \cite{CRDT98}.
In this paper, we consider a model electrostatic free boundary problem where the interface can develop regions of extremely high curvature. We develop a matched asymptotics method for conformal maps, that completely avoids the computational difficulties associated with crowding for {\em a priori unknown domains}.
An interesting interplay between interface geometry and induced stresses exists in electromechanical systems. When a conducting
interface is deflected by
an external field, the induced charge distribution
/the local electrostatic pressure is higher in the parts of the sheet with high curvature. This in turn causes even higher deflections and sharper curvatures. This can produce extremely sharp and nearly singular shapes, as well as {\em the pull-in instability}, a runaway effect due to the positive feedback between curvature and electrostatic forces, whereby no equilibrium solutions exist beyond a critical forcing strength. For these reasons we choose an electromechanical model as a test case for applying our techniques for multi-scale conformal maps. This model also has connections to recent studies that have been primarily driven by the development of microelectromechanical systems (MEMS), including micro-scale capacitors and actuators \cite{Bernstein2000,Chuang2010}.
The paper is organized as follows: In section~\ref{sec:model} we present the basic electromechanical model, a two fluid system where the interface is governed by a balance between electrostatic forces, gravity and surface tension. In section~\ref{sec:conformal} we present a reformulation of our model problem in terms of conformal mappings, and discuss the advantages of such a reformulation. In section~\ref{sec:collocation} we present the basic numerical method for constructing numerical conformal maps, and apply it to situations where the free boundary is not ``sharp". The method breaks down as the interface develops sharp features, and in section~\ref{sec:multi_scale} we present our matching technique for conformal maps that applies in a regime where the free boundary is sharp. We present a concluding discussion in section~\ref{sec:discussion}.
\subsection{Electromechanical model} \label{sec:model}
\begin{figure}[!ht]
\includegraphics[width=0.9 \linewidth]{images/model.pdf}
\caption{\label{fig:generic_model}Generic model cross-section. The interface is between a dielectric and conducting fluid, and is gravitationally stable. $\mathbf{e}_x$ is in the horizontal direction, $\mathbf{e}_y$ is vertically upwards, and $\mathbf{e}_z$ is out of the page. The acceleration due to gravity is $-g \mathbf{e}_y$.}
\end{figure}
We consider a two fluid system -- an infinite conducting fluid initially in the half-space $y < 0$, and an infinite dielectric fluid in the half space $y > 0$. Gravity acts in the $-\mathbf{e}_y$ direction, and the density of the upper fluid $\rho_0$ is less than the density of the lower fluid $\rho_0 + \rho$, so the interface is gravitationally stable. The interface is deflected by a line charge with strength $q$ per unit length, along the $\mathbf{e}_z$ direction, placed in the upper fluid at a distance $l$ above the undeflected interface. Since all the induced charge on the fluid will reside on its surface, we can equally well consider a model with an infinite conducting sheet on an elastic foundation. The conducting sheet is initially at $y=0$ and is subsequently pulled up by a line charge placed at $y=l$. The restoring force from the foundation can be modelled as a linear function of the deflection from $y =0$, and this agrees with the expression for the hydrostatic pressure on the interface in the two-fluid model, which is also a linear function of the deflection. The setup is depicted in Fig.~\ref{fig:generic_model}, and by translation invariance in the $z$-direction, it suffices to consider a two dimensional problem in the $xy$ plane with a one dimensional interface $y = h(x)$.
The pressure balance along the deflected interface $y = h(x)$ is
\begin{equation*}
-\sigma \kappa(x) + \rho g h(x) = \frac{\epsilon_0}{2} \left(\frac{\partial \phi}{\partial n}\right)^2 \text{ for } x \in (-\infty,\infty).
\end{equation*}
where
$\sigma$ is the surface tension of the interface, $\displaystyle{\kappa(x) = \frac{h_{xx}(x)}{(1+h_x^2(x))^{3/2}}}$ is the curvature of the interface at $(x,h(x))$,
$\epsilon_0$ is the dielectric permittivity of the upper fluid, $\mathbf{n}$ is the unit normal to the interface, and $\displaystyle{\frac{\partial \phi}{\partial n} = \nabla \phi \cdot \mathbf{n} = |\nabla \phi|}$ is the magnitude of the electric field on the conducting interface. When forced by a line charge, the system is closed by the equations
\begin{align*}
\nabla^2 \phi(x,y) & = -\frac{q}{\epsilon_0}\delta(x,y-l) \text{ in } \lbrace h(x) < y : x \in (-\infty,\infty) \rbrace, \\
\phi(x,h(x)) & = 0 \text{ for } x \in (-\infty,\infty).
\end{align*}
The electric field $\nabla \phi$ vanishes as $x \to \pm \infty$, so we get $h(x) \to 0$ as $x \to \pm \infty$, {\em i.e} gravity stabilizes the interface for large $x$.
In order to simplify the equations, we pick units to make $\sigma = \rho g = \epsilon_0/2 = 1$. This is equivalent to the non-dimensionalization: $$
x' = \frac{x}{l_c}, y' = \frac{y}{l_c}, h' = \frac{h}{l_c}, l' = \frac{l}{l_c},
$$
for the lengths where $l_c = \sqrt{\sigma/\rho g}$ is the capillary length, and
$$
q' = \sqrt{\frac{\epsilon^3_0}{2}}\left(\frac{\rho g}{\sigma^3}\right)^{1/4}q, \phi' = \sqrt{\frac{\epsilon_0}{2}}\left(\frac{\rho g}{\sigma^3}\right)^{1/4} \phi.
$$
We will henceforth work exclusively with the non-dimensionalized equations so there is no confusion in dropping the primes. This yields the nondimensional equations
\begin{subequations} \label{e:h_gov_pt}
\begin{align}
-\kappa(x) + h(x) & = |\nabla \phi(x,h)|^2 \text{ for } x \in (-\infty,\infty), \label{e:h_gov_pt_1} \\
\nabla^2 \phi(x,y) & = -q\delta(x,y-l) \text{ in } \lbrace y \geq h(x) \rbrace , \label{e:h_gov_pt_2} \\
\phi(x,h(x)) & = 0 \text{ for } x \in (-\infty,\infty) \label{e:h_gov_pt_3} \\
h(x) \rightarrow 0 & \; \text{as} \; x \rightarrow \pm \infty.
\end{align}
\end{subequations}
The system has a reflection symmetry $x \to -x$, and for simplicity we will restrict our attention to symmetric solutions satisfying $h(-x) = h(x), \phi(-x,y) = \phi(x,y)$. The system has two dimensionless parameters, a non-dimensional charge per unit length (i.e forcing) $q$ and a non-dimensional aspect ratio $l$, the ratio of the vertical length scale (the charge location) and the horizontal length scale, the capillary length. The local electrostatic pressure $\vert \nabla \phi \vert^2$ at each point on the sheet depends on the entire deflection profile $h(x)$ through the boundary conditions on $\phi$. This non-local coupling prevents the direct calculation of closed form solutions of \eqref{e:h_gov_pt}, and a numerical treatment is required instead.
\section{Conformal mapping reformulation} \label{sec:conformal}
We will now reformulate the system \eqref{e:h_gov_pt} in terms of conformal mappings, with the aim being to ``simplify" the non-local nature of the problem. $w = u+iv$ will denote the complex coordinate in the preimage plane and $z = x+iy$ the coordinate in the physical plane.
Assume that we have a smooth interface $y = h(x)$ with $h(x) \to 0$ as $|x| \to \infty$. Then the upper fluid domain $\Omega_u = \{y > h(x)\}$ is a proper subset of $\mathbb{C}$. The Riemann Mapping theorem \cite{conway-book} thus asserts the existence of a bijective conformal map $z=F(w)$ such that $i F(D) = \Omega_u$, where $D$ is the (open) unit disk $|w| < 1$ in the preimage $w$-plane. This mapping is unique if we additionally impose the ``normalization" $i F(0) = i l$ and symmetry $F(\overline{w}) = \overline{F(w)}$ where the overbar denotes complex conjugation. The symmetry along with the unboundedness of $\Omega_u$ implies the ``boundary condition" $\lim_{w \to -1} |F(w)| = \infty$. The mapping $F$ is illustrated schematically in figure~\ref{fig:mapping}.
\begin{figure*}[!ht]
\begin{center}
\subfigure[]
{
\includegraphics[width=0.45 \linewidth]{images/pt_full_symmetry_demonstration2.pdf}
\label{f:pt_full_symm_demo2}
}
\subfigure[]
{
\includegraphics[width=0.45 \linewidth]{images/pt_full_symmetry_demonstration1.pdf}
\label{f:pt_full_symm_demo1}
}
\caption[Schematic diagrams showing candidate deflection profiles produced using conformal maps with appropriate symmetry conditions]{\label{fig:mapping} \subref{f:pt_full_symm_demo2} Example candidate deflection profile $z = x+iy = \lbrace if(\theta) : \theta \in (-\pi,\pi) \rbrace$ given by \eqref{eq:representation}. The deflection is symmetric about $x = 0$ and decays towards $y = 0$ as $\theta \rightarrow \pm \pi$. \subref{f:pt_full_symm_demo1} Unit circle in the preimage ($w = u+iv$) plane.
}
\end{center}
\end{figure*}
For the undeflected interface $h(x) = 0$, it is easy to check that the appropriate conformal mapping \cite{nehari-conformal} is
$$
F_0(w) = l \frac{1-w}{1+w}
$$
For a deflected interface, assuming it is smooth and converges to the undeflected interface as $x \to \pm \infty$, we show rigorously (mathematical details in a subsequent paper) that the conformal map $F$ is a ``bounded" modification of $F_0$ {\em i.e.}
\begin{equation}
F(w) = \alpha \frac{1-w}{1+w} + B(w)
\label{eq:representation}
\end{equation}
where $B$ is analytic on the {\em closed} unit disk $|w| \leq 1$, $\alpha$ is real, and we have the requirements
\begin{subequations}
\begin{align}
B(\overline{w}) & = \overline{B(w)} & &\mbox{Symmetry} \label{eq:symmetry}\\
\alpha + B(0) & = l & & \mbox{Normalization} \label{eq:norm}\\
B(-1) & = 0 && \mbox{Boundary condition} \label{eq:bc}
\end{align}
\end{subequations}
The representation \eqref{eq:representation} shows that conformal map $F$ is continuous on $D \cup (\partial D \setminus\{-1\})$, i.e the only singularity of the map on the closed unit disk $|w| \leq 1$ is a pole at $w = -1$ with a {\em real} residue $2 \alpha$ corresponding to the boundary conditions $h(x) \to 0$ as $x \to \pm \infty$. The interface is then defined in terms of the conformal map $F$ parametrically by taking limits from inside the disk, i.e.:
$$
z = i f(\theta) \equiv \lim_{w_j \in D, w_j \to e^{i \theta}} i F(w) \mbox{ for }\theta \in (-\pi,\pi).
$$
This yields $z = i f(\theta) = \alpha \tan(\theta/2) + i B(e^{i \theta})$, a {\em conformal parameterization} of the the interface.
The invariance of Laplace's equation under conformal transformations now allows us to solve \eqref{e:h_gov_pt_2} using the electrostatic potential in the preimage plane to obtain
$$
\phi(z) = \frac{q}{2 \pi} \log|F^{-1}(z)|.
$$
A direct calculation using the parameterization $z = if(\theta) = x(\theta) + i h(x(\theta))$ so that $f(\theta) = h(x(\theta))-ix(\theta)$ shows that
\begin{align*}
h(x(\theta)) & = \text{Re}[f(\theta)] \\
\kappa(x(\theta)) & = \frac{x'(\theta) h''(\theta) - h'(\theta) x''(\theta)}{(x'(\theta)^2 + h'(\theta)^2)^{3/2}} = \frac{\text{Im}[f_{\theta\theta}(\theta)\overline{f_{\theta}}(\theta)]}{\vert f_{\theta}(\theta) \vert^3}
\end{align*}
The force balance \eqref{e:h_gov_pt_1} thus reduces to a single equation for $f$:
\begin{equation}
\frac{q^2}{4 \pi^2 \vert f_{\theta}(\theta) \vert^2} - \text{Re}[f(\theta)] + \frac{\text{Im}[f_{\theta\theta}(\theta)\overline{f_{\theta}}(\theta)]}{\vert f_{\theta}(\theta) \vert^3} = 0. \label{eq:new_ode}
\end{equation}
Observe that this equation is {\em local} in $\theta$ unlike the original formulation \eqref{e:h_gov_pt} where the force balance equation is {\em nonlocal}. Also observe that we have one ``real" equation for a ``complex" function $f$, so we should expect additional condition(s) that determine $f$.
For any function $\Upsilon$ that is analytic on a domain containing the closed unit disk, the real and imaginary parts of $\upsilon(\theta) = \Upsilon(e^{i \theta})$ are not independent. Rather, they are related by the periodic Hilbert transform \cite{garling-inequalities}
\begin{align*}
\text{Re}[\upsilon] & = -\mathcal{H}[\text{Im}[\upsilon]] + \text{Re}[\Upsilon[0]] \\
\text{Im}[\upsilon] & = \mathcal{H}[\text{Re}[\upsilon]] + \text{Im}[\Upsilon[0]]
\end{align*}
because they are boundary values of conjugate Harmonic functions on the unit disk \cite{conway-book}. The periodic Hilbert transform $\mathcal{H}$ of a periodic function $\xi$ is defined \cite[\S 11.5]{garling-inequalities} by
$$
\mathcal{H}[\xi](\phi) = \lim_{\epsilon \to 0} \frac{1}{2\pi}\int_{\epsilon}^\pi \cot\left(\frac{\theta}{2}\right) \left[\xi(\phi-\theta) -\xi(\phi+\theta)\right] d\theta
$$
We will rewrite this condition in a slightly different form that exploits the natural mapping between analytic functions on the closed unit disk and Fourier series given by
$$
\Upsilon(w) = \sum_n a_n w^n \leftrightarrow
\sum_n a_n e^{i n \theta} = \upsilon(\theta)
$$
where the sums do indeed converge because the coefficients $a_n$ decay exponentially, and $\upsilon(\theta) = \Upsilon(e^{i \theta})$. Since $\Upsilon$ is analytic at 0, it follows that
\begin{equation}
a_{-n} = \frac{1}{2 \pi} \int_0^{2 \pi} \upsilon(\theta) e^{i n \theta} d \theta = 0 \mbox{ for } n = 1,2,\ldots
\label{eq:constraint}
\end{equation}
These equations are {\em local} in the transform domain (in terms of $n$), but are {\em nonlocal} integral equations in terms of $\theta$. From \eqref{eq:representation}, it follows that $\Upsilon(w) = (1+w) F(w)$ is analytic on the closed unit disk, so choosing
$$
\upsilon(\theta) = \Upsilon(e^{i \theta}) = (1+e^{i \theta}) f(\theta),
$$
in \eqref{eq:constraint} yields
\begin{equation}
\int_0^{2 \pi} (1+e^{i \theta}) f(\theta) e^{i n \theta} d \theta = 0 \mbox{ for } n = 1,2,\ldots
\label{eq:hilbert}
\end{equation}
The symmetry condition \eqref{eq:symmetry} implies
\begin{equation}
f(-\theta) = \overline{f(\theta)} \mbox{ for } \theta \in (-\pi,\pi)
\label{f_symmetry}
\end{equation}
Finally, we remark that the normalization \eqref{eq:norm} is {\em nonlocal} in terms of $f$ since
\begin{equation}
l = F(0) = \Upsilon(0) = \frac{1}{2 \pi} \int_0^{2 \pi} (1+e^{i \theta})f(\theta) d \theta,
\label{f_norm}
\end{equation}
and the boundary condition \eqref{eq:bc} can be recast as
\begin{equation}
\lim_{\theta \to \pm \pi} f(\theta) = 0 \mp i \infty \label{f_bc}
\end{equation}
Equations \eqref{eq:new_ode},\eqref{eq:hilbert}--\eqref{f_bc} together give a reformulation of \eqref{e:h_gov_pt} in terms of a (normalized) conformal parameterization of the interface. The key advantage of this reformulation is that, every equation is local either on the unit circle in the preimage plane (i.e. in terms of $\theta$) or in the Fourier transform domain (i.e. in terms of $n$). Since we can efficiently switch between functions on the unit circle and their Fourier transforms using FFTs \cite{num-recipies}, this reformulation, if discretized appropriately, will give an efficient numerical algorithm for solving \eqref{e:h_gov_pt}.
\section{Collocation method}
\label{sec:collocation}
To solve for the interface numerically, we will discretize the representation for $F$ in \eqref{eq:representation}. For clarity we use a tilde to denote discretized numerical approximations, {\em e.g.} $\widetilde{F}$ is a numerical approximation to $F$. Since every analytic function on the unit disk has a convergent Taylor series \cite{conway-book}, we can define families of conformal maps with finitely many degrees of freedom by truncating the power series for $B$ centered at 0:
\begin{equation}
\widetilde{F}(w) = \alpha \left(\frac{1-w}{1+w}\right) + \sum_{j=0}^{M} \beta_j w^j.
\label{eq:discretize}
\end{equation}
In our numerics we always take $M = 2^K$, a power of 2. Then $\widetilde{f}(\theta)$ given by restricting $\widetilde{F}$ to the unit circle automatically satisfies \eqref{eq:hilbert}. Explicitly
\begin{equation}
\widetilde{f}(\theta) = -i \alpha \tan \left(\frac{\theta}{2}\right) + \sum_{j=0}^{M} \beta_j e^{i j \theta} \mbox{ for } \theta \in (-\pi,\pi). \label{eq:full_maps}
\end{equation}
The symmetry requirement \eqref{f_symmetry} implies that $\alpha, \beta_0,\beta_1,\ldots,\beta_{M}$ are all real, so the family of conformal maps in \eqref{eq:full_maps} satisfying \eqref{eq:hilbert} and \eqref{f_symmetry} is parameterized by $M+2$ real quantities. The normalization \eqref{eq:norm} yields
\begin{equation}
l = \alpha + \beta_0 \label{at0}
\end{equation}
and the boundary condition \eqref{eq:bc} implies that
\begin{equation}
\sum_{j=0}^M (-1)^j \beta_j = 0 \label{at-1}
\end{equation}
We now turn to discretizing \eqref{eq:new_ode}. Let us first consider the problem of determining $\widetilde{f}$ given $q$ and $l$. In general, we cannot expect the (numerical) residual
\begin{equation}
\displaystyle{R[\widetilde{f}] \equiv \frac{q^2}{4 \pi^2 \vert \widetilde{f}_{\theta}(\theta) \vert^2} - \text{Re}[\widetilde{f}(\theta)] + \frac{\text{Im}[\widetilde{f}_{\theta\theta}(\theta)\overline{\widetilde{f}_{\theta}}(\theta)]}{\vert \widetilde{f}_{\theta}(\theta) \vert^3}}
\label{eq:residual}
\end{equation} to vanish for all $\theta \in (-\pi,\pi)$ for our numerical maps with finitely many degrees of freedom.
The residuals at points $\theta$ and $-\theta$ are {\em not independent} since $\widetilde{f}(-\theta) = \overline{\widetilde{f}(\theta)}$ so that $R[\widetilde{f}](\theta) = R[\widetilde{f}](-\theta)$. Consequently, we get a possible discretization of \eqref{eq:new_ode} for every choice of $M$ distinct points $\lbrace\theta_m:m=0,1,\ldots,M-1\rbrace $ such that $\theta_m \neq -\theta_n$ for any pair $m \neq n$, Without loss of generality, we can take these $M$ points in $[0,\pi)$, i.e with non-negative angles. As we discussed in Sec.~\ref{sec:conformal}, we will get efficient numerical algorithms if we can easily switch between values of the function $f(\theta_m)$ and its Fourier coefficients $a_n$. This can be achieved using FFTs if we pick the points $\theta_m$ uniformly spaced on the unit circle and $M = 2^K$. In particular, for the discretization in \eqref{eq:full_maps}, we can apply the FFT to compute the function values and derivatives at evenly spaced nodes $\theta_m = m\pi/M$, $m = 0,\ldots,M-1$.
The quantities $\alpha$ and $\beta_j, j = 0,1,2,\ldots,M$ are now determined from \eqref{at0},~\eqref{at-1}~and the conditions~$R[\tilde{f}](\theta_m) = 0$ for $m=0,1,2,\ldots,M-1$. In implementing this method, we find that the numerical algorithm is better conditioned if we instead parameterize the maps by the $M+2$ quantities $\alpha$ and $h_m \equiv h(\theta_m) = \text{Re}[f(\theta_m)]$.In this parameterization the boundary condition \eqref{eq:bc} implies that $h_M = 0$, so we have $M+1$ independent real parameters. We also have $M+1$ real conditions from the vanishing of the residual at $\theta_m, m=0,1,2,\ldots,M-1$ and the normalization \eqref{at0}. The coefficients $\beta_j$ and the ``height" parameters $h_0,h_1,\ldots,h_{M-1}$ are of course related by a linear transformation which can be expressed in terms of the Discrete Fourier transform.
The final point to consider is that we {\em do not expect} the system \eqref{e:h_gov_pt} will have a unique global solution for all given $q$ and $l$, since we expect the system to display multiple equilibrium solutions for some choices of $q$ and $l$ based on the analogy with the MEMS model \cite{Pelesko2002}. As in the analysis of the MEMS model, we use a continuation method \cite{Pelesko2002} with the tip height of the interface, $h_0$ and the charge location $l$ as the specified parameters. $l$ and $h_0$ are thus the independent quantities, and the equilibrium equations determine the charge $q$ and the other heights $h_1,h_2,\ldots,h_{M-1}$ as functions of $h_0$ and $l$. Given $l$ and $h_0$, we compute the residual $R[\widetilde{f}](\theta_m)$ as a function of $(q,h_1,h_2,\ldots,h_{M-1})$ through the following steps:
\begin{enumerate}
\item Compute the Fourier coefficients $\beta_j$ corresponding to the height parameters $h_m$ where $h_M = 0$.
\item Compute $\widetilde{f}(\theta_m),\widetilde{f}_\theta(\theta_m)$ and $\widetilde{f}_{\theta\theta}(\theta_m)$ using the Inverse FFT and the representations $\widetilde{f}(\theta) = \sum \beta_j e^{i j \theta}, \widetilde{f}_\theta(\theta) = \sum i j \beta_j e^{i j \theta}$ and $\widetilde{f}_{\theta \theta}(\theta) = \sum - j^2 \beta_j e^{i j \theta}$.
\item Compute the residual $R[\widetilde{f}]$ at the points $\theta_m, m = 0,1,2,\ldots,M-1$ using \eqref{eq:residual}.
\end{enumerate}
We then compute a numerical approximation of the true deflection profile and corresponding charge strength $q$ by minimizing $\sum_m (R[\widetilde{f}](\theta_m))^2$ using a nonlinear least squares method implemented in the MATLAB function {\tt lsqnonlin} \cite{matlab}. Note that every (local) minimum for the sum of the squares of the residual is {\em not necessarily an approximate solution} to \eqref{e:h_gov_pt}. We in fact need that the {\em residual vector be zero} to within a prescribed tolerance.
Figure \ref{fig:orig_bifn_diags} shows examples of bifurcation diagrams computed using this method for $l \in \lbrace 0.25,0.5,0.75,1 \rbrace$. For all these choices of $l$, the $h(0)$ vs. $q$ curves exhibit saddle-node bifurcations i.e. they ``turn around" at a critical value of $q^*(l)$. For values of $q > q^*(l)$, the system \eqref{e:h_gov_pt} has {\em no solution}. Physically, if we perform the experiment of increasing $q$ quasi-statically from 0 past the critical value $q^*$, the tip height $h_0(q)$ will track the lower branch of the appropriate curve in Figure~\ref{fig:orig_bifn_diags}, until $q = q^*$ at which point the conducting fluid will be drawn to the point charge ``grounding" it. This is the the analog in the two-fluid system of the pull-in instability at finite forcing strength $q$ as reported for small aspect ratio MEMS \cite{Bernstein2000}.
Deflections below and above the saddle-node bifurcation are stable and unstable respectively depending on the sign of $\displaystyle{\frac{d h_0}{d q}}$. Indeed, along the upper branch, a small perturbation of $h_0$ away from the equilibrium value $h_0(q)$ will result in an interface that is subject to a higher forcing than necessary to support it (if the deflection is toward the charge) or a lower forcing than necessary (if the deflection is away from the charge). In either case, the interface will evolve dynamically and either be drawn to the point charge or relax to the (stable) equilibrium on the lower branch, so the upper branch of equilibria is unstable to small perturbations.
The solid dots on the figure represent the largest $h_0$ (for each $l$) for which the method converges with $M = 256$. While this limiting $h_0$ increases with $M$, it does so very slowly and we cannot get much closer to the vertical axis for any value of $M \lesssim 4096$ (the bound on $M$ reflects our limitations on memory size and processor speed). Nonetheless, the figure is very suggestive that, for all $l$, the bifurcation curve approaches the point $(q=0,h_0=l)$ (the solid dots on the vertical axis).
\section{Multiple scale analysis} \label{sec:multi_scale}
Figure~\ref{fig:orig_solutions} shows the numerically computed solutions for different tip heights $h_0$ with $l$ fixed at 1. The figure shows both stable ($h_0 \lesssim 0.45$) and unstable profiles. The highest profile that we can compute before the collocation method breaks down with $M = 256$ is $h_0 \approx 0.96$.
\begin{figure*}[!ht]
\begin{center}
\subfigure
{
\includegraphics[width=0.45 \linewidth]{images/bif_diags.pdf}
\label{fig:orig_bifn_diags}
}
\subfigure
{
\includegraphics[width=0.45 \linewidth]{images/solutions-eps-converted-to.pdf}
\label{fig:orig_solutions}
}
\caption{\subref{fig:orig_bifn_diags} Computed bifurcation diagrams (sheet tip height $h(0)$ plotted against charge strength $q$) for charge heights $l$ = 0.25, 0.5, 0.75, 1 (left to right). All the curves have a maximum value of $q = q^*(l)$ beyond which there are no solutions. This indicates the existence of a maximally-deflected stable solution at a critical forcing strength, numerically demonstrating a {\em saddle-node} bifurcation for this system. The upper branch of the bifurcation curves correspond to (dynamically) unstable profiles because $h_0'(q) < 0$. There is also a (independent) critical deflection magnitude, the solid dots at the end of each curve, beyond which the numerical scheme used breaks down due to crowding. $M = 256$ for all the curves. The solid dots on the vertical axis represent the charge locations, and it is suggestive that, if continued, the bifurcation curves will converge to these points. \subref{fig:orig_solutions}
Numerically computed solutions $h(x)$ with the charge location $l=1$ and $h_0$ varied. Each profile corresponds to a single solid dot on the $l=1$ (rightmost) curve in \subref{fig:orig_bifn_diags}. The solutions were generated using the collocation method with $M=256$. The dashed curves and are unstable profiles and the solid curves below are stable. For $l=1$, the maximally deflected profile that we can compute with the collocation method corresponds to $h_0 = 0.96$. }
\end{center}
\end{figure*}
It is evident from the figure that the tips (region near $x = 0$) are getting very sharp on the scale of the capillary length (nondimensionalized to 1) as $h_0$ approaches $l$. This reflects {\em non-uniformity/multiple scale behavior} of the solutions of \eqref{e:h_gov_pt}. In particular, the small scale structures are governed by a different balance (surface tension vs electrostatic pressure) from the large scale structure (surface tension vs gravity). A natural approach that will accounts for this non-uniformity it to discretize the problem {\em adaptively}, i.e. in a manner which reflects the local length scale in the solution. One would like to have roughly equal number of points to resolve each of the regions with a different dominant balance.
A weakness of the conformal mapping method, as described in Section~\ref{sec:collocation} is that it is intrinsically {\em non adaptive}. We do not get to pick the points $\theta_m$ which are the ``nodes" in our discretization of the continuous system. Rather, they are required to be uniformly spaced on the preimage plane in order to use FFT techniques. Figure~\ref{f:nodes}
shows the locations in the physical plane (i.e. on the interface) of uniformly space nodes on the unit circle for $l = 1, h_0 = 0.96, M = 256$. Contrary to our initial expectation, the breakdown of the collocation method is not due to poor resolution of the tip; rather it is due to poor resolution of the decaying region of the interface, where the dominant ``large-scale" balance is between gravity and surface tension, and the relevant length scale for this region is the capillary length.
\begin{figure}[!ht]
\includegraphics[width=0.9 \linewidth]{images/pt_full_l1_N256_solutions_with_nodes.pdf}
\caption{\label{f:nodes} Deflection profile computed with $l = 1$, $h(0) = 0.96$, $M = 256$. The density of the image nodes $\lbrace i f(i j \pi/2M) : j = 0,\ldots,M-1 \rbrace$ decreases rapidly away from the interface tip. This leaves the decaying portion of the interface poorly resolved. The bulk of the nodes are in the near-tip region which is extremely well-resolved.}
\end{figure}
This effect is easily understood by recognizing that the electric field on the interface is related to the inter-node separation in the physical plane since $|\nabla \phi| \sim 1/|f_\theta| \sim (\text{node separation})^{-1}$. Since the electric field (and the induced charge) concentrates near sharp tips, the nodes in the physical plane will concentrate near sharp tips. The tendency for image nodes to accumulate near regions of relatively high interface curvature, known as \emph{crowding}, is ubiquitous when using conformal maps defined on the unit disk $D$ to represent interfaces that feature very disparate length scales. It has been recognized as a substantial impediment to using numerical methods to find conformal maps for regions that have multiple length-scales/singularities in their boundary \cite{Wegmann2005,Porter2005}.
This is not a peculiarity of the conformal mapping based method: alternative methods that we could have applied, including boundary element methods based on the boundary integral formulations, or direct minimization of the total system energy, must all be adapted to deal with the disparate length scales that characterize stiff problems. Another approach to avoid crowding is to pick the nodes adaptively (and possibly non-uniformly) in the preimage plane. In this case, FFT methods do not apply. The Hilbert transform constraint relating the real and the imaginary parts of the interface parameterization $f$ yields a linear system with a matrix that has a poor condition number \cite{Wilkening2011}. This constraint has to be imposed along with the minimization of the residual \eqref{eq:residual}, and Wilkening \cite{Wilkening2011} has applied this idea with careful and sophisticated numerical methods to analyze the crests of large amplitude standing water waves, which were conjectured to have a corner singularity in the free interface \cite{PenneyPrice1952}. In section~\ref{sec:matching} below, we detail a different approach to this problem, one that borrows from the philosophy of matched asymptotics in multiple scale analysis \cite{Hinch1991}.
\subsection{Concentrated charge approximation} \label{sec:concentrated}
The collocation method breaks down when the tip of the interface $h_0$ approaches the charge location $l$. This motivates the definition $\epsilon = l-h_0$, as the relevant small parameter that defines this asymptotic regime. When $\epsilon$ is small, we expect that the tip is sharp and the total charge $-q$ induced on the sheet accumulates (and hence, image nodes crowd) near the high-curvature sheet tip.
As $\epsilon \to 0$, all the charge concentrates on the tip, so to leading order in $\epsilon$, we can model the charge on the interface as a concentrated charge $-q$ at the tip $(0,h_0)$.
The interface is then effectively supported by an upward electrostatic point force of strength
\begin{equation}
\Lambda \approx q^2/2\pi(l-h_0) \label{force}
\end{equation}
applied to the tip, and relaxes under the influences of gravity and elasticity everywhere else (where the electric field/induced charge is zero in this leading order approximation). The resulting leading order deflection profiles are governed by a collapsed form of the equation \eqref{e:h_gov_pt_1},
\begin{equation}
\Lambda \delta(x) - h_{out}(x) + \frac{h''_{out}(x)}{(1+(h'_{out}(x))^2)^{3/2}} = 0, \label{elastic}
\end{equation}
which can be integrated exactly.
Multiplying \eqref{elastic} by $h_{out}'$ and integrating in from infinity yields
\begin{equation}
\frac{1}{\sqrt{1 + (h'_{out}(x))^2}} + \frac{(h_{out}(x))^2}{2} = 1, \label{e:pt_out_ode_1st_int}
\end{equation}
where the constant term on the right hand side has been determined by applying the asymptotic boundary condition. We can solve this equation, provided that $0 \leq h_{out}(x) < \sqrt{2}$, to get the implicit representation
\begin{widetext}
\begin{equation}
\log{\left(\frac{2+\sqrt{4-(h_{out}(x))^2}}{2+\sqrt{4-(h_{out}(0))^2}}\right)} - \log{\left(\frac{h_{out}(x)}{h_{out}(0)}\right)} + \sqrt{4-(h_{out}(0))^2} - \sqrt{4-(h_{out}(x))^2} = |x|
\label{eq:outer-profile}
\end{equation}
\end{widetext}
For $0 \leq h_{out}(x) \leq h_{out}(0) < \sqrt{2}$, we see that the first, third and fourth terms in the above expression are bounded and $O(1)$. Consequently, for large $|x|$, we have
\begin{equation}
h_{out}(x) \simeq h_{out}(0)e^{-|x|}
\label{eq:outer-outer-profile}
\end{equation}
Integrating \eqref{elastic} across $x = 0$, we see that $h_{out}$ is continuous but has a corner (discontinuous slope) at $x=0$:
\begin{equation}
\frac{h'_{out}(0^+)}{\sqrt{1+(h'_{out}(0^+))^2}}-\frac{h'_{out}(0^-)}{\sqrt{1+(h'_{out}(0^-))^2}} = - \Lambda,
\label{eq:jump}
\end{equation}
where $\Lambda$ is the total force on the interface from the charge (or equivalently, by Newton's third law, the force on the charge pulling it towards the interface). This equation is a statement of the balance between the surface tension forces on a corner and the total electrostatic force on the interface.
We will call the solutions in \eqref{eq:outer-profile} the leading order, large deflection solutions, because they are the appropriate limit solutions as $\epsilon \to 0$ (tip curvature $\to \infty$) and they cannot be obtained as a perturbation expansion in powers of $q$ about the $q=0, h = 0$ undeformed interface. In particular, they are not smooth and have a corner at $x=0$, unlike the perturbation series solutions in powers of $q$ which are necessarily smooth. We define a parameter $\eta=\eta(h_{out}(0))$ such that $\pi/\eta(h_{out}(0))$ denotes the outside angle of the leading order interface tip corner, so that $1/2 \leq \eta \leq 1$ (see Fig.~\ref{fig:schematic_solns}). Eq.~\eqref{e:pt_out_ode_1st_int}
now yields
\begin{equation}
\eta = \left(2 - \frac{2}{\pi}\arctan{\left( \frac{2-(h_{out}(0))^2}{h_{out}(0)\sqrt{4-(h_{out}(0))^2}} \right)}\right)^{-1},
\label{eq:gamma}
\end{equation}
where $\arctan$ takes values in $[0,\pi/2]$.
Figure~\ref{fig:schematic_solns} shows the comparison between the solutions in \eqref{eq:outer-profile} and the numerically computed profile $h(x)$ from the collocation method with $l=1$. Note that the solutions agree very well outside the tip region, but in the tip region, there are consistent deviations between the leading order, large deflection solution and the numerical solution of the interface. Electrostatic forces are important near the tip, and our crude approximation of the electrostatic force is insufficient to resolve the tip region.
An important point to note is that $h_{out}(0) \neq h(0) = h_0$. This brings up the question of how to determine the appropriate value of $h_{out}(0)$ that corresponds to a given $h_0$ with the charge location fixed at $l$. We address this question in more detail below, but the basic idea is that the ``right" parameters to describe the sharp-tip asymptotic regime are $h_{out}(0)$ and $\epsilon \equiv l - h_0$. In terms of these parameters, $h(0) = h_0$ and $l$ are given by asymptotic expansions $h_0 = h_{out}(0) + m_1 \epsilon + o(\epsilon)$, $l = h_{out}(0) + n_1 \epsilon + o(\epsilon)$ where $n_1 = m_1 +1$ depends on $h_{out}(0)$ but not on $\epsilon$.
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=\linewidth]{images/compare_out_exact-eps-converted-to.pdf}
\end{center}
\caption[Leading order large deflection profiles plotted against full deflection profiles] {\label{fig:schematic_solns}Leading order large deflection profiles ($h_{out}(0) = 0.985$, $h_{out}(0) = 0.65$, dashed curves) corresponding to computed interfaces from Fig.~\ref{fig:orig_solutions} ($h_0 = 0.97$. $h_0 = 0.47$, solid curves). The inset shows a zoomed view near the point $(0,1)$. The width of the region in which the leading order large deflection approximation is inaccurate approaches zero as $\epsilon \rightarrow 0$ and the outer tip angle is $\displaystyle{\frac{\pi}{\eta}}$.}
\end{figure}
Combining Eqs.~\eqref{force},~\eqref{e:pt_out_ode_1st_int}~and~\eqref{eq:jump}, we obtain
$$
q \approx \sqrt{\left(2\pi h_{out}(0) \sqrt{4-(h_{out}(0))^2}\right)(l-h_0)}
$$
for fixed $l$. In order to compare this expression with the numerical result in Fig.~\ref{fig:orig_bifn_diags}, we need to compute (or estimate) $h_{out}(0)$ for a given $h_0$ and $l$. As we argue above
$h_{out}(0) - h_0$
is $O(\epsilon)$, so we get
\begin{align}
q & = \sqrt{\left(2\pi h_0 \sqrt{4-h_0^2}\right)(l-h_0)} + O(\epsilon^{3/2}) \label{leading_order_bifn_relationship}
\end{align}
This relationship fully describes the leading order bifurcation diagram behavior $q \sim \sqrt{l-h_0}$ in the small-$\epsilon$ regime where numerical methods face the greatest challenges. Figure \ref{f:pt_leading_h_vs_q_small_l} compares a typical computed bifurcation diagram from Fig.~\ref{fig:orig_bifn_diags} to the corresponding leading order large deflection diagram based on \eqref{leading_order_bifn_relationship}. The leading order approximation is seen to provide a consistent extension of the computed bifurcation diagram to the small $\epsilon$ regime and captures the overall bifurcation structure of the system even for $O(1)$ values of $\epsilon$.
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=0.9 \linewidth]{images/pt_full_N256_bifn_diags_with_large_defl_approx.pdf}
\caption[Example bifurcation diagrams produced using the collocation method of \S\ref{--S:elec_forward-colloc} compared to those produced using the leading order large deflection approximation of \S\ref{--S:elec_match-pt-outer}]{\label{f:pt_leading_h_vs_q_small_l} The bifurcation diagrams of Fig.~\ref{fig:orig_bifn_diags} (solid curves) compared to diagrams based on leading order large deflection profiles (dashed curves) with $l_0 \in \lbrace 0.25, 0.5, 0.75, 1 \rbrace$. The leading order solutions give consistent extensions of the computed bifurcation diagrams to small-$\epsilon$ solutions, as well as capture the overall saddle-node bifurcation structure of the system. Unsurprisingly, the accuracy is generally poor when $\epsilon/l$ is $O(1)$.}
\end{center}
\end{figure}
\subsection{Outer conformal maps} \label{sec:outer}
We now outline a new matching scheme based on numerical conformal mappings that combines high accuracy for all values of $\epsilon$ with efficient handling of the small $\epsilon$ regime. The scheme modifies leading order large deflection profiles with a more subtle description of the true tip shape. We adopt the usual notation of matched asymptotics from this point forward, labeling the conformal maps that describe leading order large deflection profiles `outer' solutions, and the conformal maps that represent the sheet tip profile `inner' solutions.
We remark that rather than just the solution profile $h_{out}(x)$ (see \eqref{eq:outer-profile}) we need an appropriate conformal parameterization $ig(\psi) = x(\psi) + i h_{out}(x(\psi))$ for the outer region. The use of conformal maps to represent both inner and outer solutions is essential for maintaining the fundamental relationship between global sheet geometry and electric field that characterizes this problem, {\em i.e.} the localization of the force balance in \eqref{eq:new_ode}.
In the same complex-variable framework as in the original numerical scheme, outer solutions are governed by the equation
\begin{equation}
- \text{Re}[g(\psi)] + \frac{\text{Im}[g_{\psi\psi}(\psi)\overline{g_{\psi}}(\psi)]}{\vert g_{\psi}(\psi) \vert^3} = 0, \label{outer_ode}
\end{equation}
where $\lbrace i g(\psi): \psi \in (-\pi,\pi]\rbrace$ is the corresponding deflection profile. This is obtained from \eqref{eq:new_ode} by dropping the electrostatic pressure term, and it is indeed Eq.~\eqref{elastic} away from $x=0$.
The relevant quantities in \eqref{outer_ode}, namely the deflection of the interface and its curvature, are properties of the profile $h_{out}(x)$ but are independent of the particular parameterization $g(\psi)$ of the interface, with the restriction that $\psi = 0$ is the tip and $\psi \to \pm \pi$ correspond to the boundary condition $h_{out}(x) \to 0$ as $x \to \pm \infty$. Thus, if a conformal map $G$ on $D$ has boundary values $g(\psi)$ which satisfies \eqref{outer_ode}, then so will the composite map $g\circ \mathcal{A}$ where $\mathcal{A}(w)$ is any conformal automorphism from the closed unit disc $D \cup \partial D$ to itself \cite[Chapter 8]{SteinShakarchi} that fixes the points $w=1$ and $w=-1$. Since all conformal automorphisms defined on $D \cup \partial D$ also map the unit circle $\partial D$ to itself, these compositions alter the parametrization but not the shape of the original deflection profile. This symmetry of the equation is akin to ``gauge freedom" in other physical systems, where one is allowed to make certain transformations which keep all the ``physical" quantities (here the deflection profile) invariant. The collection of conformal automorphisms of $D \cup \partial D$ fixing $1$ and $-1$ forms a one parameter group of mappings
$$
\mathcal{A}_a(w) = \frac{w-a}{1-aw} \text{ for }a \in (-1,1).
$$
with the group composition law
$$
\mathcal{A}_a \circ \mathcal{A}_b = \mathcal{A}_c \text{ where } \frac{1-a}{1+a} \cdot \frac{1-b}{1+b} = \frac{1-c}{1+c}.
$$
We use a collocation method, as in sec~\ref{sec:collocation} to compute a numerical conformal map that (approximately) solve \eqref{outer_ode}. An important consideration in developing a collocation method to solve for a conformal map, is to first construct a finite dimensional family of conformal maps, that is {\em adapted} to the particular problem we are trying to solve. In particular, the family of maps in \eqref{eq:discretize} are appropriate for interfaces which decay to zero and are otherwise smooth. This family is therefore {\em not appropriate} for seeking solutions of the outer equation~\eqref{outer_ode}. While the outer profile $h_{out}(x)$ still decays as $x \to \pm \infty$, the profile is no longer smooth -- It has a corner with an outer angle $\pi/\eta$ at $x = 0$.
In order to incorporate this corner with a know angle, we consider a family of conformal maps given by the composition of the family of smooth maps in~\eqref{eq:discretize} with a fixed analytic map which generates a corner.
We define the ``corner map" $C_{\eta}$ by
\begin{equation}
C_{\eta}(\zeta) = \frac{\zeta^{\frac{1}{\eta}}}{(\zeta+t)^{\frac{1}{\eta} -1}} + h_{out}(0)
\label{corner-map}
\end{equation}
where $t > 0$ is an arbitrary parameter. This map is analytic and single valued except on a branch cut which we can choose to be segment $[-t,0]$. Also, $C_\eta(0) = h_{out}(0)$, and the image of the imaginary axis under the map $C_\eta$ will have a corner at $h_{out}(0)$ with the correct ``outer" tip angle $\pi/\eta(h_{out}(0))$. Finally, we emphasize that the parameter $t > 0$ is arbitrary and a useful numerical check of our method is to verify that our results are independent of the specific choice we make for $t$.
As we discuss in Appendix~\ref{apndx:outer}, we get an appropriate family of candidate maps for approximating the outer solution by composing this corner map with the family of maps in \eqref{eq:discretize}. In particular, we seek (approximate) solutions of the outer equation within the family of conformal maps
\begin{equation}
\widetilde{G}(w) = C_\eta\left(\alpha \frac{1-w}{1+w} + \sum_{j=0}^{M_{out}} \beta_j w^j - h_{out}(0) -\frac{t(\eta-1)}{\eta}\right)
\label{outer-conformal}
\end{equation}
We numerically determine the constants $\beta_j$ and $\alpha$ that give the outer solution by requiring that the function $\widetilde{g}(\psi) = \widetilde{G}(e^{i \psi})$ satisfies the appropriate normalization and boundary conditions, and \eqref{outer_ode} at the {\em collocation points} $\psi_m = \frac{m \pi}{M_{out}}, m=0,1,2,\ldots,M_{out}-1$. The details of our numerical implementation are presented in the Appendix~\ref{apndx:outer}.
\subsection{Inner conformal maps} \label{sec:inner}
The sharp corners of outer solutions are regularized by elastic tension close to the tip as shown in Fig.~\ref{fig:schematic_solns}. Electrostatic pressure and sheet curvature are both high near such geometry, so we anticipate the existence of a scaling in which the stabilizing effect of gravity is relatively small. As evident from Fig.~\ref{fig:schematic_solns} and from the fact that the outer solutions give a corner at $x = 0$, the break down of the outer solution occurs on a length scale $\epsilon$ that is set by the separation between the charge location and the interface tip. Thus we set $f = h_{out}(0) + \epsilon \gamma$ (equivalent to scaling all physical lengths by $\epsilon$ but recognizing that the tip is within $O(\epsilon)$ of the prescribed $h_{out}(0)$) and $q = \sqrt{\epsilon} Q$ (motivated by the scaling in \eqref{leading_order_bifn_relationship}) in \eqref{eq:new_ode}. Dropping terms that are $O(\epsilon)$ results in a balance between the electrostatic pressure and surface tension:
\begin{equation}
\frac{Q^2}{4 \pi^2 \vert \gamma_{\theta}(\theta) \vert^2} + \frac{\text{Im}[\gamma_{\theta\theta}(\theta)\overline{\gamma_{\theta}}(\theta)]}{\vert \gamma_{\theta}(\theta) \vert^3} = 0 \label{inner_ode}
\end{equation}
where $\gamma$ and $Q$ are both $O(1)$, and $\lbrace i \gamma(\theta): \theta \in (-\pi,\pi)\rbrace$ is the corresponding rescaled deflection profile near the sheet tip. This equation, along with the condition $\gamma(-\theta) = \overline{\gamma(\theta)}$ has two one-parameter symmetries. We can translate the solutions by a real constant $\gamma \to \gamma+c$, because the equation only sees derivatives of $\gamma$, and we have not as yet imposed any boundary conditions on $\gamma$ which break this translation invariance. The second symmetry is the scaling $\gamma \to k \gamma, Q \to \sqrt{k} Q$ for all $k > 0$. This is related to physical observation that without gravity (and hence without the capillary length scale), the equation should have a scale invariance, as it has no intrinsic length scale. As in the previous section, we therefore do not expect unique solutions to \eqref{inner_ode} without specifying additional {\em normalization(s)} of the solution. And with an appropriate normalization, we expect to find a one parameter family of solutions that are determined entirely by the specified value of $0 < h_{out}(0) <\sqrt{2}$ or equivalently $\eta(h_{out}(0))$, as this was the only input in determining the normalized outer solution.
We now discuss the appropriate normalization conditions as well as the discretization of \eqref{inner_ode} in order to numerically compute the inner solutions which describe the tip region. An immediate observation from the form of \eqref{inner_ode}, or \eqref{eq:new_ode} with gravity neglected, is that the curvature is non-positive everywhere. In order to match the outer solutions (which have a corner at 0), it is necessary that as $\theta \to \pm \pi$, the inner profile $\gamma(\theta)$ should approach a pair of straight line asymptotes, that make an (outer) angle $\pi/\eta$. This situation is depicted schematically in Fig.~\ref{fig:inner_schematic}.
\begin{figure}[h!]
\includegraphics[width = 0.9 \linewidth]{images/inner_schematic-eps-converted-to.pdf}
\caption[Schematic inner solution]{\label{fig:inner_schematic} Schematic rescaled inner solution map $\Gamma(\omega)$. The maps depend on an arbitrary ``normalization" parameter $T$. A check for our numerical method is that as $T$ is varied, the curves should shift by a corresponding amount in the $x$-direction. This graph is rotated relative to the profiles $(x,h(x))$ since the profile is represented by $ih_0 + i \epsilon \Gamma$. }
\end{figure}
Let $\Gamma$ denote the conformal map on the (open) unit disk $D$ whose boundary values on the unit circle determine $\gamma$. We normalize away the scale invariance for $\Gamma$ by setting the separation between the tip of the interface and the charge location at $1$, and incorporate the value of $\eta$ by setting ``boundary conditions" on the behavior of $\Gamma(\omega)$ as $\omega \to -1$:
\begin{subequations} \label{e:pt_in_bcs_recast}
\begin{align}
\Gamma(0) - \Gamma(1) & = 1, \label{e:pt_in_bcs_recast_1} \\
\Gamma(\omega) -\left[C_{\text{asy}} + \left(A\frac{1-\omega}{1+\omega}\right)^{\frac{1}{\eta}}\right]& \to 0 \text{ as } \omega \to -1 \label{e:pt_in_bcs_recast_2}
\end{align}
\end{subequations}
where $C_{\text{asy}}$ and $A$ are constants whose values are determined as part of the numerical solution. We refer to $C_{\text{asy}}$ as the `asymptotic offset' of a computed solution. We still have the translation symmetry $\Gamma \to \Gamma+ c, C_{\text{asy}} \rightarrow C_{\text{asy}}+c, A \to A, Q \to Q$ for any $c \in \mathbb{R}$.
For our numerical solutions, we need a family of conformal maps that is adapted to the boundary conditions for the inner solution. As we show in Appendix~\ref{apndx:inner}, an appropriate family of conformal maps for the inner solutions is
\begin{equation}
\widetilde{\Gamma}(\omega) = \left[A \frac{1-\omega}{1+\omega} + \sum_{j=0}^{M_{in}-1} C_j \omega^j + C_{M_{in}} \left(\frac{1+\omega}{2}\right)^{\frac{1}{\eta}-1}\right]^{\frac{1}{\eta}}
\label{inner-conformal}
\end{equation}
that depends on $M_{in}+2$ real parameters $C_j, j=0,1,2, \ldots,M_{in}$ and $A$. To get a unique solution, we break the translation invariance of the solutions by specifying the additional requirement
\begin{equation}
\Gamma(1) = T \text{ for some } T > 0. \tag{\ref{e:pt_in_bcs_recast}c} \label{e:pt_in_bcs_recast_3}
\end{equation}
The constant $T$ is (somewhat) arbitrary and the restriction $T > 0$ is explained in the appendix. The values $\beta_j$ and $\alpha$ that give the outer solution by requiring that the function $\widetilde{\gamma}(\theta) = \widetilde{\Gamma}(e^{i \theta})$ satisfies the appropriate normalisation and boundary conditions, and \eqref{inner_ode} at the {\em collocation points} $\theta_m = \frac{m \pi}{M_{in}}, m=0,1,2,\ldots,M_{in}-1$. (Details in Appendix~\ref{apndx:inner}).
For intermediate values of $\eta$ (not too close to $1/2$ or equivalently $h_{out}(0)$ not close to $\sqrt{2}$) and $T$ moderately close to $0$, the algorithm for computing inner solutions produces curves with all of the desired properties (Fig.~\ref{f:pt_in_comp_prof}).
\begin{figure*}[!ht]
\begin{center}
\subfigure[]
{
\includegraphics[width = 0.45 \linewidth]{images/pt_in_T-0_9-3_gamma0_6_profs.pdf}
\label{f:pt_in_comp_prof}
}
\subfigure
{
\includegraphics[width = 0.45 \linewidth]{images/asymptotic_shift-eps-converted-to.pdf}
\label{f:pt_in_offset_vs_T}
}
\caption[Computed inner profiles]{\label{fig:inner_profile} \subref{f:pt_in_comp_prof} Computed inner solution profiles for $\eta = 0.6$, $0.05 \leq T \leq 4$. The profile for $T = 0.05$ is clearly inaccurate; the profiles for $T \gtrapprox 0$ less obviously so. \subref{f:pt_in_offset_vs_T} Asymptotic offsets for the profiles in figure~\ref{f:pt_in_comp_prof} (solid curve), plotted against $T$. The dashed curve shows the theoretical relationship \eqref{e:pt_in_Casy_T_relationship}.}
\end{center}
\end{figure*}
A check of our numerical method is that the physical results should be independent of the specific choice of $T$ (as with the choice of the parameter $t$ in computing the outer solutions). For instance, the asymptotic offset $C_{\text{asy}}$ can be computed in terms of the coefficients in the map $\widetilde{\Gamma}$ as $C_{\text{asy}} = \frac{1}{\eta}C_{M_{in}} A^{\frac{1}{\eta}-1}$. The symmetries of the equation now imply that we should have the identity
\begin{equation}
C_{\text{asy}}(T) = \frac{1}{\eta}C_{M_{in}} A^{\frac{1}{\eta}-1} = T + k \label{e:pt_in_Casy_T_relationship}
\end{equation}
for some constant $k$ depending only on $\eta$ and independent of $T$. This relation is not respected by computed solutions if $T$ becomes either very close to $0$ or larger than a (\emph{a priori} unknown) threshold value (Fig.~\ref{f:pt_in_offset_vs_T}). For the value of $\eta$ used in these demonstrations, the expected linear behavior is only displayed for $0.1< T < 1$. In addition to giving a check on our numerical method, computing profiles for several $T$ values allows us to identify the range that produces consistent solutions.
\subsection{Matching: Uniformly valid composite conformal maps} \label{sec:matching}
Now that we can compute inner and outer solutions separately, we need to combine them to produce an approximate solution of the full system \eqref{eq:new_ode}. This is not straightforward, as conformal maps are ``rigid" global objects. Conformal maps on the disk are necessarily smooth (infinitely differentiable), and the real and the imaginary parts must satisfy the nonlocal relation \eqref{eq:constraint}. Consequently, we cannot simply patch together two conformal maps on different portions of the disk, to produce a composite conformal map. In fact, if two conformal maps agree in a small neighborhood of a single point, then by analytic continuation they necessarily are identical, i.e. they have a common maximal domain of analyticity and the maps are identical on this domain \cite{Ahlfors-book}. Since the inner and the outer conformal maps are not identically equal (for instance near the tip), they are different ``everywhere", and any map that agrees with one cannot agree with the other. At first sight one might conclude there {\em there is no way to put two distinct conformal maps together}.
Of course, we do not require that the composite map agree {\em exactly} with either the inner or the outer conformal map. Rather the combined map should approximate the inner and the outer conformal maps ``well" on different pieces of the unit disk. We will now turn this qualitative notion into a quantitative algorithm for combining inner and outer conformal maps to generate a composite map to solve \eqref{eq:new_ode} with {\em a uniformly small error}.
As discussed above the outer equation \eqref{outer_ode} and the inner equation \eqref{inner_ode} have nontrivial symmetries. In order to compute (unique) numerical solutions of these equations, we break these symmetries by imposing additional normalization conditions. We now have to ``un-normalize" i.e., pick the appropriate symmetry transformed versions of the computed inner and the outer solutions, so that, together they give a solution for the original system \eqref{eq:new_ode}.
This is achieved by the {\em matching principle}, which requires that there exist an overlap region where both the inner and outer solution describe the true physical solution, and thus agree with each other {\em asymptotically} \cite{vanDyke1975,Hinch1991}.
The full (symmetry related) family of conformal maps describing the inner solution is given by $\widetilde{F}(\omega) = h_{out}(0) + \epsilon(\widetilde{\Gamma}(\omega) + c)$, where $\omega \in D$ is the {\em inner variable} and $\epsilon > 0,c \in \mathbb{R}$ are elements of the symmetry group, describing scaling and vertical translation of the inner solution. The ``outer limit" of this inner solution is the behavior of this map as $\omega \to -1$ and from our representation for the inner solution a direct calculation yields (see also \eqref{e:pt_in_bcs_recast_2})
\begin{align*}
\widetilde{F}_{in}(\omega) = & h_{out}(0) + \epsilon \left[\left(A\frac{1-\omega}{1+\omega}\right)^{\frac{1}{\eta}} +\frac{C_{M_{in}} A^{\frac{1}{\eta}-1}}{\eta}+c\right] \\
& + \epsilon O((1+\omega)^{\frac{1}{\eta}}).
\end{align*}
The full (symmetry related) family of conformal maps describing the outer solution is given by $\widetilde{F}_{out}(w) = \widetilde{G}(w)$, where the reparameterization invariance of the outer solution implies that the {\em outer variable} $w$ has no intrinsic physical meaning unless it is $1$ (corresponding to the tip) or $-1$ corresponding to the point at $\infty$. Rather if $w$ and $w'$ are two variables related by a conformal automorphism that fixes $\pm 1$,
$$
w' = \mathcal{A}_a(w) = \frac{w-a}{1-aw}, \quad a \in (-1,1)
$$
then both $\widetilde{G}(w') = \widetilde{G
}\circ \mathcal{A}_a(w)$ and $\widetilde{G}(w)$ are valid descriptions of the outer solution.
In contrast, the inner variable $\omega$ does have intrinsic physical meaning ($\omega = 0$ at the charge location, for instance). Thus, fixing the symmetry element in the reparametrization group is the same as determining $a \in (-1,1)$ such that the inner solution $\widetilde{F}_{in}(\omega)$ matches with the outer solution $\widetilde{F}_{out}(\omega) = \widetilde{G}(\mathcal{A}_a(\omega))$ in an overlap region. We write this relation in shorthand as ``the" outer variable $w = \mathcal{A}_a(\omega)$, meaning the outer variable is identified with that particular choice of $w$ for which the matched outer solution is given by the normalized map $\widetilde{G}$ computed numerically as in section~\ref{sec:outer}.
A direct calculation reveals that for $w = \mathcal{A}_a(\omega)$
$$
\frac{1-w}{1+w} = \frac{1+a}{1-a} \cdot \frac{1-\omega}{1+\omega}.
$$
so to match the inner and outer solutions, it is convenient to represent the outer limit of the inner solution and the inner limit of the outer solution in terms of $(1-\omega)/(1+\omega)$ and $(1-w)/(1+w)$ respectively.
Van Dyke's matching rule \cite{Hinch1991} requires that the outer limit of $\widetilde{F}_{in}$ agree with the inner limit of $\widetilde{F}_{out}$. We can implement this rule by expanding $\widetilde{F}_{in}$ in $(1+\omega)$, $\widetilde{F}_{out}$ in $(1-w)$ and using the relation $w = \mathcal{A}_a(\omega)$ between $w$ and $\omega$. This yields the conditions
\begin{align}
c & = - \frac{C_{M_{in}} A^{\frac{1}{\eta}-1}}{\eta} \label{eq:matching_1} \\
\epsilon^\eta A & = t^\eta \frac{\left(\alpha - 2 \sum_{j=0}^{M_{out}}j \beta_j\right)}{t} \cdot \frac{1+a}{1-a} \label{eq:matching_2}
\end{align}
The quantities $A,\alpha,\beta_j$ and $C_{M_{in}}$ are defined in \eqref{outer-conformal} and \eqref{inner-conformal}. They are obtained by solving problems \eqref{outer_ode} and \eqref{inner_ode} that are {\em independent} of $\epsilon$. Consequently they are $O(1)$ quantities which can only depend on $h_{out}(0)$ and not on $\epsilon$. Matching thus gives the requirement
$$
\frac{1+a}{1-a} = \left(\frac{\epsilon}{t}\right)^\eta \frac{At}{\left(\alpha - 2 \sum_{j=0}^{M_{out}}j \beta_j\right)} \equiv K \epsilon^\eta
$$
where we have defined an $O(1)$ constant $K$. The overlap function is calculated using \eqref{eq:matching_1} in $\widetilde{F}_{in}$ to be
\begin{align}
\widetilde{F}_{c}(\omega) = & h_{out}(0) + \epsilon \left(A\frac{1-\omega}{1+\omega}\right)^{\frac{1}{\eta}} \nonumber \\
= & h_{out}(0) +t \left(\frac{\alpha - 2 \sum_{j=0}^{M_{out}}j \beta_j}{t}\right)^{\frac{1}{\eta}}\left(\frac{1-w}{1+w}\right)^{\frac{1}{\eta}}, \label{eq:overlap}
\end{align}
From the success in applying Van Dyke's matching rule it follows that
\begin{enumerate}
\item In the ``outer" region $|1+\omega| \ll 1$, the inner solution agrees very well with the overlap function,
\item In the ``inner" region $|1-w| \ll 1$, the outer solution agrees very well with the overlap function,
\item There is a non-trivial overlap region where $|1-w| \ll 1$ and $|1+\omega| \ll 1$. In this region, all the three functions $\widetilde{F}_{in}, \widetilde{F}_{out}$ and $\widetilde{F}_{c}$ agree with each other,
\end{enumerate}
where by ``agree well" we mean that the leading terms in an asymptotic expansion in $\epsilon \to 0$ are identical. This argument can also be recast in terms of {\em matching through an intermediate variable} \cite{Hinch1991}. For any given $0 < r < \eta$ and complex number $\tau$ with $\text{Re}(\tau) > 0$, there is a sufficiently small $\epsilon_0$ such that for all $\epsilon < \epsilon_0$, $-1+\epsilon^r \tau \in D$, the open unit disk. From the relation \eqref{eq:matching_2} between $w$ and $\omega$, we get
\begin{align}
& \text{Re}(\tau) > 0, \quad
\omega = -1+\epsilon^r \tau \nonumber \\
\implies & \quad w = \mathcal{A}_a(\omega) = 1-\frac{4 K \epsilon^{\eta-r}}{\tau} + O(\epsilon^\eta).
\label{eq:intermediate}
\end{align}
Substituting these expressions in $\widetilde{F}_{in}$ and $\widetilde{F}_{out}$ respectively, we get
\begin{align*}
\widetilde{F}_{in} & = h_{out}(0) + \epsilon^{1-\frac{r}{\eta}} \left(\frac{2A}{\tau}\right)^{\frac{1}{\eta}} + O(\epsilon^{1-\frac{r}{\eta} + 2 r})\\
\widetilde{F}_{out} & = h_{out}(0) + \epsilon^{1-\frac{r}{\eta}} \left(\frac{2A}{\tau}\right)^{\frac{1}{\eta}} + O(\epsilon^{1-\frac{r}{\eta} + (\eta- r)})\\
\end{align*}
so the first two terms in the asymptotic expansions of the inner and outer conformal maps agree as $\epsilon \to 0$ with $\tau$ fixed.
Using $\omega \sim O(1)$ and $\omega \sim -1 + \epsilon^{\eta} \tau$ as the ``boundaries" of the overlap region, we determine that the inner and the outer profiles agree in ``physical" space for $\epsilon \ll x \ll 1$. This makes good physical sense because the the overlap region should be all the length scales between the tip curvature scale where electrostatics balances curvature and the capillary length where gravity balances curvature. In this intermediate regime curvature dominates both gravity and electrostatics, so the solutions correspond to $\kappa = 0$, i.e. straight lines. Figure~\ref{f:pt_map_overlap} shows the overlap region, described by straight lines, where the inner and outer solutions agree.
\begin{figure*}[!ht]
\begin{center}
\subfigure[]
{
\includegraphics[height=2.5 in]{images/overlap_error.pdf}
\label{f:pt_mat_overlap_region1}
}
\subfigure[]
{
\includegraphics[height=2.5 in]{images/overlap_error_small_eps.pdf}
\label{f:pt_mat_overlap_region2}
}
\label{f:pt_mat_overlap}
\caption[Plots showing the overlap regions for two inner/outer profile pairs]{\label{f:pt_map_overlap} Inner and outer solution profiles (solid lines) for $l_1 = 1$, $M_{out} = 128$, $M_{in} = 32$. Each profile is compared to the overlap profile given by the dashed line. By comparing the profiles computed using \subref{f:pt_mat_overlap_region1} $\epsilon = 0.04$ and \subref{f:pt_mat_overlap_region2} $\epsilon = 0.01$, we observe that the region in which both inner and outer profiles agree with the overlap profile to within a given tolerance is of the form $\epsilon k < x < K$ where $k$ and $K$ are both $O(1)$.}
\end{center}
\end{figure*}
We can now construct a composite numerical conformal map $\widetilde{F}$ on $D$ which is a uniformly valid approximation \cite{Hinch1991} (i.e works both in the inner and the outer region) by
\begin{equation}
\widetilde{F}(\omega) = \widetilde{F}_{in}(\omega) + \widetilde{F}_{out}(\omega)-\widetilde{F}_{c}(\omega).
\label{eq:composite}
\end{equation}
This map, as a linear combination of conformal maps, is manifestly a conformal map on the unit disk. For $|1+\omega|$ small, i.e. in the outer region $\widetilde{F} \approx \widetilde{F}_{out}$ and for $|1-w|$ small, i.e. in the inner region $\widetilde{F} \approx \widetilde{F}_{in}$. This composite conformal map therefore approaches the right ``limits" asymptotically in the inner and the outer regions (which overlap) and is thus an approximate solution with a uniformly small error. The interface profile is obtained by taking the limit $\widetilde{f}(\theta) = \lim_{\omega \to e^{i \theta}} \widetilde{F}(\omega)$.
Our matching procedure is schematically illustrated in figure.~\ref{fig:two-circles}. Since we construct the composite map by solving for the inner and the outer conformal maps separately, the underlying domain on which the numerical conformal map $\widetilde{F}$ is defined is naturally interpreted as a {\em pair of (discretized) unit circles}. The outer and the inner variables are related by $w = \mathcal{A}_a(\omega)$ and the matched profile in the (physical) $z$-plane is given by $z = \widetilde{F}(\omega)$ as in \eqref{eq:composite}. Note the non-uniformity in the relative distribution of the inner an the outer nodes. We emphasize that, for matching conformal maps, the overlap region, as described by the intermediate variable in \eqref{eq:intermediate} is {\em two dimensional}. In particular, our matching procedure requires that the inner and the outer solutions agree on a open set in the complex plane, and it does not suffice, in general, to only have agreement of the inner and the outer profiles, i.e the functions representing $h(x)$.
On the unit disks in the $\omega$- and $w$-planes, the overlap region is given by the common region between $|1+\omega| \lesssim O(1)$ (the outer region) and $|1-w| \lesssim O(1)$ (the inner region). We indicate this region in the $\omega$- and $w$-planes by hatching in figure~\ref{fig:two-circles}. We remark that the densities of the inner and outer nodes are comparable in the overlap region but not outside it. This is to be expected as the inner and the outer solutions give equally well resolved numerical solutions in the overlap region. For the schematic plot in figure~\ref{fig:two-circles}, the densities do not differ by more than a factor of 2 in the overlap region, so there are no more than 4 and no fewer than 1 circle for every 2 asterisks and vice-versa. In contrast, the nodes outside the overlap region are overwhelmingly either circles (the inner region) or asterisks (the outer region), showing that in these regions either the inner or the outer solution is inaccurate/poorly resolved.
\begin{figure*}[hp]
\includegraphics[width = 0.95 \linewidth]{images/schematic-color-eps-converted-to.pdf}
\caption[Two circles perspective]{\label{fig:two-circles} Schematic representation of the matching procedure. Our matching procedure involves three complex planes, the ({\em inner}) $\omega$-plane, the ({\em outer}) $w$-plane and the ({\em physical}) $z$-plane. The circles $(\bigcirc)$ represent the {\em inner nodes} which are uniformly distributed on the unit circle in the (inner) $\omega$-plane, and the asterisks ({\Large \textasteriskcentered}) are the {\em outer nodes} which are uniformly distributed on the unit circle in the (outer) $w$-plane. They are drawn with different sizes in the $\omega$ and $w$ planes so they can be clearly distinguished. The {\em overlap} region described by the intermediate variable (see \eqref{eq:intermediate}), is hatched in the $\omega$- and the $w$-planes. The matched profile is plotted on semilog axes to clearly show how the inner nodes (circles) resolve the small $x$ region near the tip while the outer nodes resolve the large $x$ region.}
\end{figure*}
Because the inner and outer solutions have their own set of nodes, it suffices to have as few as 16 nodes in each region to get fully resolved matched solutions for the interface. The solution in Figure~\ref{fig:two-circles} was generated with 32 inner and 32 outer nodes. Figure\ref{f:family_matched_solns} show matched interface profiles created by combining a single pair of inner (32 node) and outer (128 node) profiles using different values of $\epsilon$. Generating profiles in this way (with fixed $h_{out}(0)$) is extremely efficient once the initial calculations required to determine the appropriate inner and outer solutions have been performed. Note that the charge location $l$ is not equal for all these interfaces; rather it varies with $\epsilon$. This is a minor drawback when seeking solutions of the forward problem. In order to compute the solutions with a specifies $l$ and $h_0$, we need to invert the (nonlinear!) relation between these parameters, and the asymptotic parameters $h_{out}(0)$ and $\epsilon$.
\begin{figure}[ht!]
\includegraphics[width = \linewidth]{images/matched_solutions-eps-converted-to.pdf}
\label{f:family_matched_solns}.
\caption[Families of matched profiles computed by combining a single pair of inner and outer profiles] {Matched profiles computed with $h_{out}(0) = 1$ and $\epsilon = 0.03$, $0.1$, $0.17$. Te dashed curve is the outer solution and the solid curves are the matched solutions. The inner and outer solutions only need to be calculated once to produce these plots. The inset shows the tip shapes for the profiles. The rescaling of the inner solution by $\epsilon$ produces the appropriate tip curvature in each case.}
\end{figure}
Observe that the matched profile $h(x) < 0$ for $x \gtrsim 1$ (a few times the capillary length) in figure~\ref{f:family_matched_solns}. This disagrees with the physical intuition that the force from the line charge will only deflect the interface upwards. In fact, we can prove rigorously that the sheet cannot deflect below the $x$-axis for the system \eqref{e:h_gov_pt}. This dip below the axis in our matched solutions, albeit very small, reveals a possible flaw in our matched solution and therefore demands an explanation.
In our outer solutions, the balance is between gravity and elasticity, and for large $x$, the profile $h_{out}(x)$ decays exponentially, as in \eqref{eq:outer-outer-profile}. For any multipole source, the electric field on an asymptotically flat conductor decays algebraically. Far away from the tip, the combination of the forcing charge and the induced charge on the tip is effectively a dipole of strength $q \epsilon \sim \epsilon^{3/2}$. Consequently, the induced charge decays no faster than $\epsilon^{3/2}x^{-3}$ for large $x$, and the electrostatic pressure decays as the square of the induced charge $\sim \epsilon^{3} x^{-6}$. Comparing this expression with the hydrostatic pressure in $h_{out}(x)$ we get a crossover when $\epsilon^{3} x^{-6} \sim e^{-x} \implies x \sim \log(\epsilon^{-1})$. Consequently, no matter how small is $\epsilon > 0$, there always exists an outer-outer region where the dominant balance is no longer between gravity and elasticity. Rather, it is between gravity and electrostatics. Further, the boundary of this outer-outer region, $x \sim \log(\epsilon^{-1})$ is not very different from the capillary length scale $x \sim O(1)$ unless $\epsilon$ is exceedingly small.
A formal calculation keeping higher orders in our composite matched solution confirms the conclusions from this heuristic argument. The outer-outer region is given by the crossover $\exp(-\epsilon^\eta/(1+ \omega)) \sim \epsilon |1+\omega|^{2-\frac{1}{\eta}}$ so that for
$$
|1+\omega| \lesssim -\frac{\epsilon^\eta}{\log(\epsilon)},
$$
it is no longer true that the presumed dominant term $\widetilde{F}_{out}(\omega)$ is larger than the ignored higher order corrections in the presumed smaller term $\widetilde{F}_{in}(\omega) -\widetilde{F}_{c}(\omega)$. In other words, the composite solution in no longer a valid asymptotic expansion in the outer-outer region, and one has to introduce a new ``layer" to resolve the non-uniformity in the expansion of the solution in powers of $\epsilon$ \cite{Hinch1991}. This argument also illustrates one of the strengths of the method of matched asymptotics -- {\em the composite solutions have in them, the information about when they are no longer valid} \cite{Hinch1991}.
As we argued above, the scaling for this outer-outer region, $|1+\omega| \sim \frac{\epsilon^\eta}{\log(\epsilon^{-1})}$ differs from the outer-scale $|1+ \omega| \sim \epsilon^\eta$ only by a logarithmic factor, so that for any realistic $\epsilon$, there is not a clear separation between the outer and the outer-outer regions. This makes the matching problem difficult, as is the case with other matched asymptotics problems that have logarithmic terms \cite{vanDyke1975,Hinch1991}. We will present a more extensive analysis of the outer-outer region and the associated matching problem in a future publication.
We also remark that the outer-outer region has a very small effect on the solution in the inner region near the tip, and on the computed bifurcation diagram using the matched solution shown in Fig.~\ref{f:pt_full_vs_lead_vs_mat_bifn}. The figure confirms that the matching technique described in this paper produces uniformly valid, high accuracy solutions (errors are $O(\epsilon)$ in the inner {\em and} the outer regions) for all values of $\epsilon$ while remaining computationally efficient (especially!) when $\epsilon$ is small.
\begin{figure}[th!]
\includegraphics[width = 0.9 \linewidth]{images/pt_full_N256_bifn_diags_with_large_defl_and_match.pdf}
\caption[Bifurcation diagrams from the original collocation and matching methods]{\label{f:pt_full_vs_lead_vs_mat_bifn}The directly computed (collocation method) and leading order bifurcation diagrams of Fig.~\ref{f:pt_leading_h_vs_q_small_l} (solid and dashed curves respectively) compared to the bifurcation diagram based on matched solutions (dotted curve) with $l_0 = 1$. Note the excellent agreement between the dotted and the solid curves, even for $O(1)$ values of $\epsilon$, despite the matching procedure relying on asymptotic expansions which depend on the smallness of $\epsilon$.}
\end{figure}
\section{Discussion} \label{sec:discussion}
In this paper, we develop an adaptive numerical conformal mapping method for efficiently computing multiple scale structures in 2 dimensional problems. The central result is that, with our method, we can use conformal mapping techniques to analyze structures that have widely disparate length scales. This is a significant improvement over existing numerical conformal mapping methods for free boundary problems, because they are essentially ``single-scale" and are plagued by the problem of crowding when applied to multiple-scale problems. As we discuss in Sec.~\ref{sec:matching}, there are good reasons why one would not even consider matched asymptotics for conformal maps, and to our knowledge, this is the first work that successfully combines numerical conformal mappings with the method of matched asymptotics.
We consider a model system is governed by three basic forces -- gravity, electrostatics and surface tension. The interface is governed by multiple dominant balances between different forcing/restoring mechanisms, so the free boundary does develop a multiple scale structure. The model problem is thus a good test bed for developing our method, although it is
not (to our knowledge) directly relevant to applications.
We reformulate the problem in terms of conformal mappings -- it converts a non-local problem into a problem that is governed by local equations, except some of the equations are in ``real" space an some are in ``reciprocal" space (i.e in the Fourier transform domain). Since the Fast Fourier transform (FFT) allows us to efficiently convert from real to reciprocal space and vice-versa, we can get efficient numerical methods for approximating solutions to our problem in terms of discretized conformal maps. This approach leads to the collocation method as described in section~\ref{sec:collocation}.
The FFT requires evenly spaced nodes (in both real and reciprocal space), so we are no longer free to choose the node locations in our discretization. This is a particular concern in multiple-scale problems, since the nodes tend to concentrate in regions of high curvature, leaving the rest of the domain under-resolved. This phenomenon, called crowding, is a significant impediment to the using conformal mapping methods for multiple-scale problems. The underlying issue, namely the ``stiffness" of the governing equations, also shows up in other methods (boundary integral/boundary element etc.), and is responsible for the challenges of numerically studying free boundary problems with widely disparate scales. In our model, we find that the numerical collocation method no longer converges to a solution once the tip curvature gets sufficiently large.
The ``obvious" physical parameters of the model are the non-dimensional charge strength $q$ and location $l$. However, the ``correct" asymptotic parameters in the sharp-tip regime are the ``uncorrected" tip height $h_{out}(0)$ and the tip-charge separation $\epsilon$. In section~\ref{sec:concentrated} we use physical arguments to determine the charge distribution on the interface and deduce the appropriate scaling for the the physical parameters in terms of the asymptotic parameters. We emphasize that the analysis in this section {\em did not use conformal maps}, and as such the ideas also extend to three dimensional versions the two-fluid or MEMS models.
Using scalings in terms of the asymptotic parameter $\epsilon$, we reduce the governing equation to an {\em outer} equation that describes the balance between gravity and surface tension, and an {\em inner} equation that describes the balance between electrostatic forces and surface tension. These {\em asymptotic equations} have non-trivial symmetries, and we determine {\em discretized conformal map solutions} of these equations with appropriate normalization using the collocation method as in section~\ref{sec:collocation}. It is important that we pick the discretization appropriately to reflect the known structure of the solutions (a corner in the interface for the outer solution, asymptote to a pair of straight lines with a known angle for the inner solution). Because each region has a single balance, we are not solving for a multiple-scale structure, and the collocation method works well (Sections~\ref{sec:outer}~and~\ref{sec:inner}).
In applying this method to other problems, it is crucial that we identify the ``correct" asymptotic parameters and scaling relations for the physical quantities, since this knowledge is required to derive the asymptotic limiting equations for the system. We then use the relevant symmetries of the asymptotic equations (The method is powerful precisely because conformal maps have a rich group of symmetries) to determine the appropriate matching conditions that allows us to combine distinct conformal maps on different scales to form a single composite ``multi-scale" conformal map.
In section~\ref{sec:matching}, we discuss this issue in detail for our model system. The key insight is that the natural domain for discretized multi-scale conformal maps is not a single complex plane. Rather, it is two (or perhaps more for other problems) sets of evenly spaced points on unit-circles, that are related by non-trivial mapping functions (conformal automorphisms of the unit disk). Each discretized unit circle resolves one scale (i.e region with one dominant balance) in the solution, and the requirement that the functions defined on the various circles agree (asymptotically) on their mutual overlaps allow us to patch them together to construct a composite multi-scale conformal map. Since each unit circle is discretized evenly, we can still exploit the efficiency of the FFT in solving the asymptotic equations. and the the morphing of the unit circle by conformal automorphisms allows us to (in effect) pick the nodes in our discretization adaptively and thus resolve structures with disparate length scales in the solution (see figure~\ref{fig:two-circles}).
This idea seems related to the Cross-ratios of Delaunay triangulation (CRDT) algorithm for computing Schwarz-Christoffel transformations for nonconvex/multi-armed polygonal domains \cite{CRDT98}. The CRDT algorithm is designed to combat crowding, and a key idea in is to use multiple maps into unit circles that are related by conformal automorphisms. These automorphisms are chosen to ``blow-up" portions of the unit circle to ensure that nodes are not crowded (locally). However, unlike our method of asymptotic matching of conformal maps, the CRDT algorithm exploits the invariance of the cross ratio \cite{Ahlfors-book} under conformal automorphisms to (implicitly) construct a composite conformal map. Further work is needed to explicate the precise connection between our multi-scale method for free boundary problems and the CRDT algorithm for given ``multi-scale" polygonal domains.
\section*{Acknowledgements}
This work was supported by the NSF through DMS grant 0807501. We are also grateful to two anonymous referees whose comments on an earlier version helped significantly improve the presentation and discussion of our results.
|
\section*{Methods}
\textbf{Experimental details.} The laser source was a linearly-polarized, single wavelength ($\lambda \sim$ 633 nm) helium-neon laser (Melles Griot) with a power of $\sim$ 10 mW which was expanded and collimated by a telescope ($f_{\textrm{L1}}$ = 15 mm and $f_{\textrm{L2}}$ = 125 mm) to approximate a plane wave. The plane wave illuminated a HoloEye Pluto spatial light modulator (SLM) ($1080 \times 1920$ pixels) which has a resolution of 8 $\mu$m and was calibrated for a wavelength of 633 nm. The SLM was addressed with holograms representing Durnin's ring-slit aperture encoded via amplitude modulation (see Supplementary Information). Two ring-slit apertures with $R1 = 179$ pixels (1432 $\mu$m), $R2 = 195$ pixels (1560 $\mu$m) and $d = 16$ pixels (128 $\mu$m) were used, each encoded with a non-canonical azimuthal phase variation of order $\ell$ and $-\ell$, respectively, so that the transmission function on the SLM was given by
\begin{equation}
t(r, \varphi) = \left\{
\begin{array}{l l}
\cos(\theta/2)\exp(i \ell \varphi) + \sin(\theta/2)\exp(-i \ell \varphi) & \quad R1-d/2 \leq r \leq R1 +d/2\\
\sin(\theta/2)\exp(i \ell \varphi) + \cos(\theta/2)\exp(-i \ell \varphi) & \quad R2-d/2 \leq r \leq R2 +d/2\\
0 & \quad \text{elsewhere}
\end{array} \right. .
\end{equation}
The Fourier transform of the field at the plane of the SLM (i.e. the superposition of two Bessel beams of orders $\ell$ and $-\ell$ with $k_{r1} = 28428.1$ $\textrm{m}^{-1}$ and $k_{r2} = 30969.2$ $\textrm{m}^{-1}$) was achieved with the use of a lens: $f_{\textrm{L3}}$ = 500 mm and magnified with a $10 \times$ objective. The $10 \times$ objective also acted as an aperture to select only the first diffraction order which was captured on a CCD camera (Point Grey fire-wire CCD).
\textbf{Petal position.} To extract the petal position from the captured camera images, the measured Bessel beams were compared to simulated beams with the same number of petals, which were adapted in spatial scale. The comparison was achieved in a quantitative manner by evaluating a two-dimensional cross-correlation coefficient when rotating the simulated patterns with respect to the measured beam. The correct angular rotation of the measured beam was found from a maximum of the correlation function. To avoid ambiguities in the form of a multiple number of correlation maxima, the simulated pattern was rotated within the interval of 0$^\circ$ to 180$^\circ / \ell$. The fidelity of the process was improved by iteratively refining the interval limits until a chosen accuracy was achieved, which was 0.1$^\circ$ in the experiments.
\textbf{Power exchange.} For each beam intensity $I$ recorded at a propagation distance $z$, the center, which was obtained from the first order moments, was surrounded by a circle, whose radius $R$ was manually adapted to enclose the petal region. From integrating the beam intensity inside and outside of this circular aperture, the power of the petal and ring region was determined, and normalized to the respective total power $P_\text{tot}$:
\begin{eqnarray}
P_\text{petal}&=&\frac{1}{P_\text{tot}}\iint_{x^2+y^2<R^2}I(z)dxdy\\
P_\text{rings}&=&\frac{1}{P_\text{tot}}\iint_{x^2+y^2>R^2}I(z)dxdy,
\end{eqnarray}
which is necessary due to the limited definition region of the Bessel beams (see Supplementary Information). The power exchange between the petal and the ring region is then identified from the difference of the maximum to the minimum of the petal power $P_\text{petal}^\text{max}-P_\text{petal}^\text{min}$.
|
\section{Introduction}
The Landau-Zener (LZ) formula gives the transition probability when a system is swept through an avoided crossing~\cite{lz,zener}. Explicitly, by introducing the {\it diabatic states} $|1\rangle$ and $|2\rangle$ and write a general state as $|\psi(t)\rangle=c_1(t)|1\rangle+c_2(t)|2\rangle$, the LZ problem solves the coupled equations ($\hbar=1$)
\begin{equation}
i\frac{\partial}{\partial t}\left[
\begin{array}{c}
c_1(t)\\
c_2(t)\end{array}\right]=\left[
\begin{array}{cc}
\lambda t & U \\
U & -\lambda t\end{array}\right]\left[
\begin{array}{c}
c_1(t)\\
c_2(t)\end{array}\right],
\end{equation}
where $\lambda$ is the sweep velocity and $U$ the coupling strength of the two diabatic states. For an initial state $|\psi(-\infty)\rangle=|1\rangle$, the probability for population transfer from the state $|1\rangle$ to the state $|2\rangle$ at $t=+\infty$ is $P_\mathrm{LZ}=\exp\left(-\Lambda\right)$ with the {\it adiabaticity parameter} $\Lambda=\frac{2\pi U^2}{\lambda}$. In the {\it adiabatic regime}, $\Lambda\ll1$, we obtain an almost complete transfer of population between the two states. This LZ formula holds only for initial conditions as the one above (or equivalently $|\psi(-\infty)\rangle=|2\rangle$) and for infinite integration times $t\in[-T,T]$; $T\rightarrow\infty$. For finite times or other initial conditions, quantum interference alters the exponential transition formula. As will be discussed in the present work, this phenomenon is especially evident in certain non-linear extensions of the above paradigm LZ model.
Various generalizations of the LZ problem have been considered in the past, especially multi-level problems~\cite{threelz1,bowtie,julienne}, many-body situations~\cite{altland,fleischhauer,lmgcrit,orso,LZoptlat}, and non-linear LZ transitions~\cite{nlz,wu,lmglz,lmg2}. It has been particularly demonstrated that for non-linear models both the exponential dependence and the smoothness of $P_\mathrm{LZ}$ can be lost due to {\it hysteresis} phenomena~\cite{nlz,wu}. Furthermore, in the adiabatic regime when $P_\mathrm{LZ}$ is smooth, the transition probability typically obeys a power-law dependence, {\it i.e.} $P_\mathrm{LZ}\sim\lambda^\nu$ for some exponent $\nu$~\cite{wu}. Such non-linear LZ problems arise in mean-field theories of quantum many-body problems~\cite{nlz,wu,lmglz,lmg2}. Using classical adiabaticity arguments, power-law dependences have also been predicted in many-body LZ problems beyond the mean-field regime~\cite{altland,fleischhauer}. All these works assume infinite integration times, or more precisely choosing an initial state $|\psi(-\infty)\rangle=|1\rangle$ (or the ground-state in the many-body/level setting). At these infinite initial times the diabatic and {\it adiabatic states} coincide and as a result, effects deriving from the interference phenomenon mentioned above will be greatly suppressed. It is not clear, however, how other more general initial states will evolve for non-linear models.
We note that in the above extended LZ models the transition is maintained by a constant coupling between the diabatic states. Thus, interaction in these models primarily adds an effective (non-linear) energy shift of the instantaneous ({\it adiabatic}) energies. In this work we consider a different scenario where the coupling is solely driven by interaction, such that turning off the interaction implies a trivial decoupled system. In particular, we analyze a {\it Lipkin-Meshkov-Glick model} (LMG)~\cite{lmg}, both at a mean-field and at a many-body level. At the mean-field level, by considering initial states as those discussed above ($|1\rangle$ or $|2\rangle$) they are decoupled and we encounter no population transfer. As a result, to stimulate any population transfer both initial, diabatic or adiabatic, states have to be populated and interferences between the two is unavoidable. In addition, in this ``interaction induced LZ model'', as will be shown, this type of interference has far more drastic influence on the dynamics than in the other LZ models. Beyond mean-field, at a full many-body level, quantum fluctuations will, however, act as a sort of `dephasing' and the interference phenomenon is not equally transparent. Like in other extended LZ model, both at the mean-field and the full quantum level we find a power-law dependence on the transition probability, but the exponentials differ in the two cases for the system sizes considered.
\section{Landau-Zener transitions}
Due to the diverging adiabatic energies of the LZ model in the asymptotic time limits, whenever more general initial states of the LZ problem are studied one encounters a mathematical controversy regarding quantum interferences. Let us briefly mention this by looking at the general solution of the LZ problem which can be expressed in terms of a scattering matrix;
\begin{equation}\label{smatrix}
\left[\begin{array}{c}
c_1(+\infty)\\
c_2(+\infty)\end{array}\right]=
\left[\begin{array}{cc}
S_1 & S_2\\
-S_2 & S_1^*\end{array}\right]
\left[\begin{array}{c}
c_1(-\infty)\\
c_2(-\infty)\end{array}\right].
\end{equation}
The matrix elements are~\cite{smel}
\begin{equation}
\begin{array}{l}
S_1=\sqrt{1-P_\mathrm{LZ}}e^{i\chi},\\ \\
S_2=P_\mathrm{LZ},
\end{array}
\end{equation}
with the phase
\begin{equation}\label{fas1}
\chi=\frac{3\pi}{4}-\arg\left[\Gamma\left(i\frac{\Lambda}{2}\right)\right]+2\Phi,
\end{equation}
where the last term is related to the (adiabatic) dynamical phase accumulated throughout the transition,
\begin{equation}\label{fas2}
\Phi=\lim_{t\rightarrow\infty}\left[\frac{\lambda}{2}t^2+\frac{\Lambda}{2}\log\left(\sqrt{2\lambda}t\right)\right]
\end{equation}
and $\Gamma(x)$ is the {\it gamma function}. Obtaining the asymptotic solution above implies studying the limits of functions when their arguments $|z|$ goes to infinity. These limits may depend on the phase of $z$, something referred to as the {\it Stokes phenomenon}~\cite{stokes}. The lines in the complex plane where the function changes character are called {\it Stokes lines} and in particular for the LZ problem the $t\rightarrow-\infty$ and the $t\rightarrow+\infty$ limits belong to different sectors divided by two such Stokes lines~\cite{kalle}.
Returning to the expressions (\ref{fas1}) and (\ref{fas2}) we have that $\Phi$ diverges in the large time limit, which means that the probability to find the system in, say, state $|1\rangle$ for an initial state $|\psi(-\infty)\rangle=\cos\theta|1\rangle+\sin\theta|2\rangle$,
\begin{equation}\label{prob}
\begin{array}{lll}
P_1 & = & \cos^2\theta\left(1-P_{LZ}\right)+\sin^2\theta P_{LZ}\\ \\
& & +\sin2\theta P_{LZ}\sqrt{1-P_{LZ}}\cos\chi,
\end{array}
\end{equation}
is ill-defined. Naturally, this is a result of looking at the asymptotic solution of the LZ problem, while for finite time sweeps the dynamical phase $\Phi$ is finite. This interference effect is well known from the theory of {\it Landau-Zener-St\"uckelberg interferometry}~\cite{stuck}: the LZ transition depends on the relative phase of the incoming state.
It should be clear that whenever the state is initialized in say $|1\rangle$ and the initial time is negative and finite the transition probability will always display some non-monotonic behaviour due to interference between the corresponding adiabatic states. As will be demonstrated in the next section, for the interaction induced LZ problem discussed in this work, the influences from this type of interference is greatly enhanced.
\section{Interaction induced Landau-Zener transitions}
The LMG model was first introduced in nuclear physics~\cite{lmg}, but have since then been shown to be of relevance for numerous other systems, including atomic condensates in double-well traps~\cite{lmg1,lmglz}, in ion traps~\cite{lmg2}, or in cavity/circuit QED~\cite{lmg3}. The LMG model can also be seen as an {\it infinite range transverse Ising model} where every spin interact equally with each other. The type of LMG model we analyze is given by
\begin{equation}\label{lmgham}
\hat{H}_\mathrm{LMG}=\lambda t\hat{S}_z-\frac{U}{\mathcal{S}}\hat{S}_x^2.
\end{equation}
Here, $\hat{S}_x$, $\hat{S}_y$ and $\hat{S}_z$ are the $SU(2)$ angular momentum operators obeying the commutation relations $[\hat{S}_\alpha,\hat{S}_\beta]=i\varepsilon_{\alpha\beta\gamma}\hat{S}_\gamma$ with $\varepsilon_{\alpha\beta\gamma}$ the fully antisymmetric {\it Levi-Civita tensor}. The LZ sweep velocity $\lambda$ is taken to be positive, and $U$, the interaction strength, is also positive meaning that we consider the ferromagnetic case. The ``classical limit'' accounts to take the spin $\mathcal{S}\rightarrow\infty$. The diabatic states are the eigenstates of the $z$-spin component, $\hat{S}_z|\mathcal{S},m\rangle_z=m|\mathcal{S},m\rangle_z$. Importantly, we note that it is the interaction term causing a coupling between these diabatic states. With the {\it Schwinger's spin-boson mapping}~\cite{sakurai}; $\hat{S}_z=\left(\hat{a}^\dagger\hat{a}-\hat{b}^\dagger\hat{b}\right)/2$, $\hat{S}^+=\hat{a}^\dagger\hat{b}$, and $\hat{S}^-=\hat{b}^\dagger\hat{a}$, it follows that in the boson representation the interaction scatters two `$a$'-particles into two `$b$'-particles or vice versa. In addition to the continuous $U(1)$ symmetry arising from conserved spin, $\left[\hat{{\bf S}}^2,\hat{H}_\mathrm{LMG}\right]=0$, the model also supports a $\mathbb{Z}_2$ parity symmetry given by $\left(\hat{S}_x,\hat{S}_y,\hat{S}_z\right)\rightarrow\left(-\hat{S}_x,-\hat{S}_y,\hat{S}_z\right)$. If $\mathcal{S}$ is an integer, the ground state at $t=-\infty$ and at $t=\infty$ has the same parity, while this parity changes when $\mathcal{S}$ is an half integer. In the following we will always assume the spin to be an integer such that the instantaneous ground state parity is preserved through the sweep.
Thinking of $t$ as a parameter, for large $|\lambda t|$ the ground state is ferromagnetic; $|\mathcal{S},\mathcal{S}\rangle_z$. For $\lambda t=0$ instead, the ground state $|\mathcal{S},\pm\mathcal{S}\rangle_x$ is doubly degenerate. In the thermodynamic limit (here equivalent to the classical limit $\mathcal{S}\rightarrow\infty$), the model is {\it quantum critical}~\cite{sachdev} with critical points at $\lambda t/U=\pm2$. The transitions are of the {\it Ising universality class} and for $|\lambda t/U|<2$ the system is in the symmetry broken phase in which the $\mathbb{Z}_2$ parity is broken. The antiferromagnetic LMG model~(\ref{lmgham}), {\it i.e.} $U<0$, is not critical but instead there is a first order transition at $\lambda t=0$ separating the two ferromagnetic states $|\mathcal{S},\pm\mathcal{S}\rangle_z$.
As a final remark, we compare the present LMG model to the otherwise frequently analyzed LMG systems, see Refs.~\cite{lmg1,lmglz,lmg2,lmg3}. In all these cases, a term $\epsilon\hat{S}_x$ is included in the Hamiltonian. Such a term breaks the $\mathbb{Z}_2$ symmetry and thereby split the ground state degeneracy and the model is no longer quantum critical. Equally important, the LZ transition occurs also for zero interaction $U=0$ in such cases. We call these models for {\it parity-broken LMG} systems.
\subsection{Mean-field analysis}
As the spin is preserved, the phase space is the $SU(2)$ {\it Bloch sphere} with radius $\mathcal{S}$. The classical, or mean-field Hamiltonian, depends therefor on the polar and azimuthal angles $\theta$ and $\phi$. The corresponding classical Hamiltonian
\begin{equation}\label{clasham}
\frac{H_\mathrm{cl}}{\mathcal{S}}=\lambda t\cos(\theta)-U\sin^2(\theta)\cos^2(\phi),
\end{equation}
gives the classical equations of motion
\begin{equation}\label{claseom}
\begin{array}{l}
\dot{\phi}=\lambda t\sin(\theta)+U\sin(2\theta)\cos^2(\phi),\\ \\
\dot{\theta}=U\sin^2(\theta)\sin(2\phi),
\end{array}
\end{equation}
with the dot representing the time-derivative. Note that the above mean-field equations are ``exact'' when the quantum state is enforced to populate a spin-coherent state $|\theta,\phi\rangle$, and in particular one would expect the accuracy of this approach to be good for spins $\mathcal{S}\gg1$. Initially, $t_i=-T$, we assume the {\it magnetization} $z\equiv\cos(\theta)\approx1$. Thus, the spin precesses around the north pole. This marks an important difference between the present model and previously studied ones; if we let $z\equiv1$ we see that at a mean-field level the dynamics is frozen, {\it i.e.} no population transfer takes place. This derives from the fact that the transitions are emerging from interaction and when the ``target mode'' is empty there are no (quantum) fluctuations stimulating a transition. Thereby, we automatically have to initialize $z\neq\pm1$ and as a result the population transfer will depend on the above discussed quantum interference occurring between the adiabatic states. Note that this is regardless of integration time - also in the limit $T\rightarrow\infty$. This is very different from other non-linear LZ models where the interaction acts as an effective energy shift rather than a coupling of diabatic states~\cite{nlz,wu}. Thus, we expect the LZ interference effect to be particularly pronounced in the present model.
\begin{figure}[h]
\centerline{\includegraphics[width=8cm]{mfimbalnew2a.eps}}
\caption{The imbalance $z$ for large times $t_f=200$ as a function of $\lambda/U$ (we actually average $z(t)$ over some periods around $t_f$ in order to avoid additional fluctuations). The initial condition is taken as $z(t_i)=\cos\theta=0.98$ and $\phi(t_i)=0$. The insets display zooms of the imbalance around the sweep velocity $\lambda/U\approx0.3$ and $\lambda/U\approx5$. The green solid line is the result from a TWA simulation with 5 $\%$ fluctuations in the initial imbalance $z(t_i)$ and fully random initial phase $\phi(t_i)$. } \label{fig2}
\end{figure}
In the following we will integrate the classical equations of motion from $t_i=-200$ to $t_f=+200$. The initial magnetization $z_{t_i}=0.98$, {\it i.e.} the initial spin is very close to pointing to the north pole. The magnetization after the LZ sweep, $z_{t_f}$, as a function of the LZ sweep velocity is displayed in Fig.~\ref{fig2}. In the adiabatic regime, typically $\lambda<1$, the lower bound of $z_{t_f}$ follows a power-law behaviour $\sim\lambda^\nu$ with $\nu=1/2$. The smallest $\lambda$ in the figure is $\lambda=0.02$ meaning that for $t_i=-200$ the lowest adiabatic state is predominantly populated. The LZ type interferences are evident throughout the parameter regime in terms of rapid oscillations where the transition is greatly suppressed. It is important to appreciate that the amplitude of these oscillations are much larger than what could be expected from the expression~(\ref{prob}); with $z=\cos\theta=0.98$ the amplitude of the LZ oscillations $\sin(2\theta)P_\mathrm{LZ}\sqrt{1-P_\mathrm{LZ}}<0.15$. We have numerically integrated the regular LZ problem with the same initial state and over the same time interval and found that the oscillations are often an order of magnitude smaller in amplitude in the regular LZ model. Furthermore, when the integration time is increased in the present model, the amplitude of oscillations grows and in the limit $t_i\rightarrow-\infty$ our numerical results suggest that the (adiabatic) transfer can be largely suppressed also for infinitely small $\lambda$'s. This is in stark contrast to the analytical result (\ref{prob}) for the regular LZ problem.
Let us look closer to the behaviour of Fig.~\ref{fig2} and especially how the LZ interference can be understood in this classical picture. For large negative $\lambda t$, the azimuthal angle $\phi$ oscillates rapidly while the polar angle $\theta$ evolves on a much longer time-scale ({\it adiabatic regime}). Put in other words, whenever $\lambda|t|\gg U$, which warrants adiabatic evolution, the classical action $I=\int_0^{2\pi} z\,d\phi$, with the integration curve along the classical phase space trajectory, stays constant (equivalently, during one classical orbit the Hamiltonian change is minimal)~\cite{ll}. In the vicinity of the crossing, $\lambda t\sim0$, there is, however, no clear separation of time-scales between the two variables and it is here that adiabaticity breaks down (also called {\it sudden} or {\it critical regime}). In the limit of adiabatic evolution, the state follows the instantaneous constant energy curves, $H_\mathrm{cl}[\theta,\phi,t]=\mathrm{constant}$. The extrema of the Hamiltonian functional give the fixed points of Eq.~(\ref{claseom}). The north and south pole on the Bloch sphere are two {\it hyperbolic fixed points} of the classical equations of motion. There are two additional ({\it elliptic}) fixed points; $(\theta_\mathrm {fp},\phi_\mathrm{fp})=\left(\arccos\left(-\lambda t/2U\right)\right),0)$ and $(\theta_\mathrm{fp},\phi_\mathrm{fp})=\left(\arccos\left(-\lambda t/2U\right)\right),\pi)$. For large times $|t|$ these coincide with the other two fixed points. For $|\lambda t|/2U\leq1$, however, they traverse the Bloch sphere along the meridians $\phi=0,\,\pi$. The four fixed points defines the (non-linear) {\it adiabatic energy curves} via $E_\mathrm{ad}(t)=H_\mathrm{cl}[\theta_\mathrm{fp},\phi_\mathrm{fp}]$, which with the above expressions become $E_\mathrm{ad}(t)=\pm\mathcal{S}\lambda t$ and $E_\mathrm{ad}(t)=-\mathcal{S}\left(U+\frac{\lambda^2t^2}{4U}\right)$.
Historically, the rapid changes in the transition probability for non-linear LZ problems has been traced back to a hysteresis effect (the adiabatic energies build up so called {\it swallow-tail loops}) which is present above some critical strength of non-linearity~\cite{wu}. Dynamically, this is explained from two fixed points `colliding' in phase space and the solution is not able to precess around a single fixed point any longer. The interferences of Fig.~\ref{fig2} can also be understood by returning to the phase space evolution. Initially, the system adiabatically encircles the north pole. At some instant, in the terminology of a {\it transcritical bifurcation}, two elliptic fixed points begin to depart from the north pole. This is the critical regime where there exist no clear separation of time scales. The solution can here `chose' between encircling the hyperbolic or elliptic fixed point as they separate in phase space. In the latter case, the system ends up with a large fraction of population centered around the south pole. Thus, the rapid variations in the population transfer again stems from a `collision' of fixed points, but this time it coincides with a critical point in the original quantum model. This is indeed the crucial difference between this model and the earlier studies; the fate of the system which is determined from which fixed point it will `follow' occurs in the critical regime while in other models the evolution can be smooth up to the `hysteresis jump'. Another way to see the difference is to note that for the parity-broken LMG system, the bifurcation is of the {\it imperfect} type. To make the picture more clear, the adiabatic energy curves, showing the transcritical bifurcation, are depicted in Fig.~\ref{fig3} as blue lines (solid lines are the stable and dashed lines the unstable classical solutions). Breaking of the parity symmetry implies opening up a gap between the two elliptic solutions. In fact, it has been shown that for the ferromagnetic ($U>0$) parity-broken LMG model a non-zero interaction $U$ increases the population transfer~\cite{nlzexp} contrary to the LMG model analyzed in this work. We see that the present model also displays swallow-tail loops, but contrary to earlier studies these additional solutions are always present for non-zero $U$ and the fixed point `collision' therefor occurs as long as $U\neq0$.
A most relevant question is if the interferences survive in the quantum case where quantum fluctuations could destabilize the classical solutions. To explore the influence of quantum fluctuations of the initial states we apply the {\it truncated Wigner approximation} (TWA)~\cite{twa} which solves the classical equations of motion for a set of initial states $(\phi_n(t_i),\theta_n(t_i))$ which are taken randomly according to the initial quantum distribution $|\Psi(\theta,\phi,t_i)|^2$. The resulting semi-classical results are obtained by averaging over the set of classical solutions, {\it i.e.} the trajectories are added incoherently meaning that any dynamical quantum interference effects are neglected. The results of a TWA simulation is presented as the green line in Fig.~\ref{fig2}. Expectedly, the initial fluctuations smears out the rapid variations in the fully classical results. Note, however, that deep in the classical regime (large $\mathcal{S}$), these fluctuations could, in principle, be made arbitrary small and the interferences should reappear.
\subsection{Full quantum analysis}
We now go beyond the classical and semi-classical approaches of the previous subsection and analyze the evolution of the full quantum system defined by the Hamiltonian~(\ref{lmgham}). One of the main objections is to explore whether the interference structure found in the transition probabilities in the classical model survives also in the quantum problem. Before presenting the results we may note that there are some earlier studies of related problems, but none has discussed the unavoidable interferences appearing in this model at a mean-field level. More precisely, driving the ferromagnetic LMG model through its critical point was analyzed in Refs.~\cite{lmgcrit}, and it was found that the non-adiabatic corrections obey a power-law dependence of the sweep velocity $\lambda$. A similar behaviour was also demonstrated in the {\it Tavis-Cummings model} describing $N$ spin-1/2 particles collectively interacting with a single boson mode~\cite{altland}. Also the LZ problem of the parity-broken LMG model has been considered~\cite{lmglz}.
\begin{figure}[h]
\centerline{\includegraphics[width=8cm]{lmgeig3.eps}}
\caption{The spectrum of the LMG Hamiltonian $\hat{H}_\mathrm{LMG}$ for various $\lambda t/U$. The spin $\mathcal{S}=10$. The dotted black and solid red curves correspond to the two different parity solutions. The blue solid and dashed curves give the classical stable and unstable solutions respectively. The splitting and recombination of the classical solutions show the transcritical bifurcations.} \label{fig3}
\end{figure}
The eigenenergies $\varepsilon_n$ of $\hat{H}_\mathrm{LMG}$ are displayed in Fig.~\ref{fig3}. In the thermodynamic limit, the critical points are at $\lambda t/U=\pm2$ for which the two parity states become degenerate. Since the spectrum is symmetric with respect to $\lambda t/U=0$ it follows that the spectrum of the anti-ferromagnetic LMG is simply $-\varepsilon_n$. This demonstrates the fact that the anti-ferromagnetic LMG is not critical but hosts a first order quantum phase transition.
When the initial ground state evolves it passes through a seam of level crossings starting at $t\approx-2U/\lambda$ and continuous until $t=2U/\lambda$. Thus, the system realizes a multi-channel Landau-Zener-St\"uckelberg interferometer. We note that this multi-level LZ crossings cannot, however, be described by the LZ {\it bow-tie model}~\cite{bt}. Interferences between the different paths (adiabatic states) could lead to final populations divided among the different diabatic states. In order to compare the amount of excitations in the present system to the LZ formula we introduce the projectors $\hat{P}_n(t)=|\psi_n\rangle\langle\psi_n|$, where $|\psi_n\rangle$ is the $n$'th instantaneous eigenstate of $\hat{H}_\mathrm{LMG}$, and define the excitation fraction as
\begin{equation}
P_\mathrm{ex}=\lim_{t\rightarrow\infty}\frac{1}{2\mathcal{S}+1}\sum_{n=0}^{2\mathcal{S}}n\langle\psi(t)|\hat{P}_{n+1}(t)|\psi(t)\rangle.
\end{equation}
Here, $|\psi(t)\rangle$ is the solution of the full time-dependent problem.
Thus, $P_\mathrm{ex}$ measures the amount of non-adiabatic excitations; $P_\mathrm{ex}=0$ corresponds to the case when only the ground state is populated while $P_\mathrm{ex}=1$ is the opposite limit of a maximally excited system. Note that $P_\mathrm{ex}$ is the mean of the final (scaled) distribution $P(n)$ of population of the various states $|\psi_n\rangle$. In the asymptotic limit $t\rightarrow+\infty$, when the diabatic and adiabatic states coincide, $\langle\hat{S}_z\rangle=\mathcal{S}(2P_\mathrm{ex}-1)$. To fully characterize the final distribution one would need all moments $\Delta^{(k)}n=\sum_nn^kP(n)$. Of particular interest is the Mandel $Q$-parameter~\cite{mandel}
\begin{equation}\label{qpara}
Q=\frac{\Delta^{(2)}n-\left(\Delta^{(1)}n\right)^2}{\Delta^{(1)}n}-1
\end{equation}
which says whether the distribution $P(n)$ is sub- ($Q<0$) or super-Poissonian ($Q>0$).
\begin{figure}[h]
\centerline{\includegraphics[width=8cm]{qexnewest2.eps}}
\caption{The average (scaled) number of excitations created during the LZ sweep for different spins $\mathcal{S}$: following the arrow 5 (blue), 12 (red), 24 (green), 50 (black), and 74 (magenta). In the adiabatic regime (inset) $P_\mathrm{ex}\sim\lambda^2$.} \label{fig4}
\end{figure}
The full time-dependent problem has been integrated from $t_i=-200$ to $t_f=+200$. We consider various spins $\mathcal{S}$ and sweep velocities $\lambda$. The results for $P_\mathrm{ex}$ are shown in Fig.~\ref{fig4}. We see in the figure that for growing spin $\mathcal{S}$ the system becomes more excited which can be understood from the increased density of states. Indeed, this is a result deriving from the critical slowing down mechanism in the vicinities of critical points. In the adiabatic and in the intermediate regimes we in particular find (numerically) that $P_\mathrm{ex}\sim\sqrt{\mathcal{S}}$.
In the adiabatic regime different power-law dependences $P_\mathrm{ex}\sim\lambda^\nu$ have been established in various types of LZ models; $\nu=3/4$ for parity-broken LMG model~\cite{wu}, $\nu=1$ for the Tavis-Cummings model~\cite{altland}, and $\nu=1/3$ (quantum regime) or $\nu=1$ (semi-classical regime) for a many-body fermionic model related to the Tavis-Cummings one~\cite{fleischhauer}. For a sweep through one of the critical point of the parity LMG it was found that $\nu=2$ deep in the adiabatic regime and $\nu=3/2$ in the intermediate regime~\cite{lmgcrit}. Such a dynamical situation is different from a full LZ sweep taken in this work where Landau-Zener-St\"uckelberg interferences can alter the excitations. Nevertheless, one may expect similar power-law dependences and this is indeed also the case as has been verified numerically. Thus, for small sweep velocities $\lambda$, $P_\mathrm{ex}\sim\lambda^2$ ({\it i.e.} $\nu=2$), and for the regime where breakdown of adiabaticity considerably sets in $P_\mathrm{ex}\sim\lambda^{3/2}$ ({\it i.e.} $\nu=3/2$) and finally in the diabatic regime we recover an exponential dependence $P_\mathrm{ex}\sim\left[1-\exp\left(-\kappa/\lambda\right)\right]$ for some $\lambda$-independent constant $\kappa$ (which is however $N$-dependent). Note that the corresponding quantum and semi-classical models display different scaling in the adiabatic regime. The inset of Fig.~\ref{fig4} display the excitations in the adiabatic regime, and we can hint that for large spins the exponential $\nu$ is actually smaller than 2 in the limit $\lambda\rightarrow0$ which could explain the discrepancy between the quantum and classical results; we can only expect agreement in the classical limit $\mathcal{S}\rightarrow\infty$.
We now return to the question raised in the beginning of this section, {\it i.e.} will the interference phenomenon discussed in the classical model survive also in the quantum case? Clearly, from Fig.~\ref{fig4} we see that the classical oscillations are absent in the quantum simulations. One may say that this is of no surprise since the oscillations were already gone in the TWA result. However, as already pointed out, the limit $\mathcal{S}\rightarrow\infty$ should reproduce the classical results - the oscillations should appear for large enough spin values. This is certainly true for the TWA case at finite times since then the quantum uncertainty can be made vanishingly small. In the quantum case, for spins as large as $\mathcal{S}=500$ we have not been able to see any signatures of the classical oscillations. Furthermore, we see that the semi-classical TWA evolution is in general more adiabatic compared to the quantum one for spins $\mathcal{S}>20$. The difference between the quantum and classical results should be ascribed the multi-level Landau-Zener-St\"uckelberg interferences. We expect that such interferences should generate large fluctuations in the distribution $P(n)$. If this is the case one should find large Mandel $Q$-parameters as $\mathcal{S}$ grows. In the quasi-adiabatic and intermediate regimes we have numerically found that $Q\sim \mathcal{S}$, while from Fig.~\ref{fig4} we can extract that in the corresponding regime $P_\mathrm{ex}\sim\sqrt{\mathcal{S}}$. So the fluctuations relative to the non-adiabatic excitations in the system $Q/P_\mathrm{ex}\sim\sqrt{\mathcal{S}}$ which agrees with the findings of~\cite{altland}. Thus, as $\mathcal{S}$ is increased a larger number of final diabatic/adiabatic states will become populated.
\section{Conclusions}\label{seccon}
The LZ problem of a LMG model where the transition is driven by particle interaction was studied. At the mean-field level it was demonstrated that interference can drastically affect the transition probabilities throughout the different parameter regimes, and most surprisingly also deep in the adiabatic regime. At the quantum level, on the other hand, quantum fluctuations tend to `dephase' the dynamics such that rapid oscillations in the transition probabilities are suppressed. The mean-field and quantum results predicted different scaling behaviour of the transition probabilities, but this could arise from the finite system sizes used for the quantum simulations.
We end by suggesting one possible system described by the parity LMG. By loading ultracold atoms into the first excited states of a two-dimensional anisotropic anharmonic trap, the quasi degeneracy of these states results in an effective spin-1/2 structure of the atoms~\cite{fernanda1}. The $\hat{S}_x^2$-term derives from the atom-atom interaction~\cite{fernanda1}. The anharmonicity of the trap prevents the atoms to decay into other energy states and by tuning the two trap frequencies the LZ sweep is realizable.
\begin{acknowledgments}
The author acknowledges Fernanda Pinheiro for discussions, and VR (Vetenskapsr\aa det) for financial help.
\end{acknowledgments}
|
\section{Introduction}
It has been recognized that there are profound connections between gravity and \mbox{thermodynamics \cite{Cocke,Bekenstein,Hawking,Davies,Unruh}}. Since then, these connections has been steadily becoming stronger. It has been shown that the entropy $S$ can be taken as the Noether charge associated with the diffeomorphism invariance of the theory \cite{Wald,Iyer}. In \cite{Jacobson}, the Einstein equation has been derived from the first law of thermodynamics. This attempt also has been investigated in modified gravity theories \cite{Eling,Elizalde,Brustein} and has been revisited in \cite{Makela}. In a wide class of spacetime, the field equations in both general relativity and Lovelock theories can be expressed as a thermodynamic identity near the horizon \cite{Padmanabhan,Paranjape,Kothawala} (see a \mbox{review \cite{Padmanabhan2010}}). In \cite{Gao2011}, the generalized Tolman--Oppenheimer--Volkoff equation is derived by using the maximum entropy principle to a charged perfect fluid, impling that there are fundamental relationships between general relativity and ordinary thermodynamics. The equations of motion for modified gravity theories, such as $F(R)$ gravity, the scalar-Gauss--Bonnet gravity, $F(\mathcal{G})$ gravity, and the non-local gravity, are equivalent to the Clausius relation in thermodynamics \cite{Bamba}. In \cite{Bracken}, the Einstein--Hilbert action can be constructed by minimizing the free energy. It was argued that the variation of the surface term evaluated on any null surface which acts a local Rindler horizon can be given a direct thermodynamic interpretation \cite{Parattu}. Gravity was explained as an entropic force caused by changes in the information associated with the positions of material bodies \cite{Verlinde}. Possible modifications and extensions to this interesting idea were \mbox{proposed \cite{Gao2010,Li,Cai,Hendi}}. All these studies were based on some assumptions, such as Unruh temperature, horizon, null surfaces, and so on. In \cite{Yang}, The thermal entropy density has been obtained for any arbitrary spacetime without assuming a temperature or a horizon, implying that gravity possesses thermal effects, or, thermal entropy density possesses effects of gravity.
Here we generalize the results in \cite{Yang} to the case of nonzero chemical potential. The results we obtained indicate that the changes of temperature, entropy, particles, and chemical potential will result in gravitational force, or gravitational force will induce changes of temperature, entropy, particles, and chemical potential.
\section{Relations between Gravity and Thermodynamics}
Both the energy density $\rho$ and the pressure $p$ play important roles in general relativity or thermodynamics. $\rho$ and $p$ are components of the stress-energy tensor in general relativity and are fundamental variables in thermodynamics. Let us begin with the first law of thermodynamics for fluids consisting of particles in curved spacetime
\begin{eqnarray}
\label{1st}
dE=TdS-pdV+\mu dN.
\end{eqnarray}
where $E$, $S$, and $N$ represent the total energy, entropy and particle number within the volume $V$, $\mu$, $T$, and $p$ are the chemical potential, the temperature, and the pressure of the perfect fluid, respectively. $dV=\sqrt{h}d^3x$ with $\sqrt{h}$ the determinant of the spatial metric. We take $c=G=1$ and use metric signature $(-,+,+,+)$ throughout this paper. Rewriting (\ref{1st}) in terms of densities
\begin{eqnarray}
\label{1ast}
d(\rho V)=Td(sV)-pdV+\mu d(nV),
\end{eqnarray}
we can easily get
\begin{eqnarray}
\label{1bst}
\rho dV+ Vd\rho=TVds+TsdV-pdV+n\mu dV+V\mu dn,
\end{eqnarray}
where $s$ is the entropy density and $n$ the particle number density. Applying Equation (\ref{1st}) to a unit volume, we have
\begin{eqnarray}
\label{1cst}
d\rho=Tds+\mu dn.
\end{eqnarray}
Combining Equations (\ref{1bst}) and (\ref{1cst}), we obtain the integrated form of Gibbs--Duhem relation \cite{Gao2001}
\begin{eqnarray}
\label{1dst}
Ts=\rho+P-\mu n.
\end{eqnarray}
For $\mu=0$, Equation (\ref{1dst}) reduces to the thermal entropy density obtained in \cite{Yang}. Now, considering the Einstein equation in \cite{Wald1984}
\begin{eqnarray}
\label{E}
R_{\mu\nu}-\frac{1}{2}g_{\mu\nu}R+\Lambda g_{\mu\nu}=8\pi T_{\mu\nu},
\end{eqnarray}
and the stress energy tensor of the perfect fluid
\begin{eqnarray}
\label{T}
T_{\mu\nu}=g_{\mu\nu}p+(\rho+p)u_{\mu}u_{\nu},
\end{eqnarray}
we obtain
\begin{eqnarray}
\label{E1}
R-4\Lambda =- 8\pi (3p-\rho).
\end{eqnarray}
Using the $3+1$ decomposition of Einstein equation, we derive \cite{Wald1984,Gourgoulhon}
\begin{eqnarray}
n^{\mu}n^{\nu}R_{\mu\nu}+\frac{1}{2}R-\Lambda =8\pi \mathcal{E},
\end{eqnarray}
where $n^{\mu}$ is the unit normal vector field to the 3 dimension hypersurfaces $\Sigma$ and $\mathcal{E}=\Gamma^2(\rho+p)-p$ with $\Gamma$ the Lorentz factor. According to the scalar Gauss relation, one can get
\begin{eqnarray}
\label{3E}
\mathcal{R}+K^2-K_{ij}K^{ij}-2\Lambda =16\pi \mathcal{E},
\end{eqnarray}
where $\mathcal{R}$ is the Ricci scalar of the 3 dimension hypersurfaces $\Sigma$, $K_{ij}$ the extrinsic curvature tensor of~$\Sigma$, and $K$ the trace of the $K_{ij}$.
Combining Equations (\ref{E1}) and (\ref{3E}), we obtain the expression of $\rho+p$ in general relativity \cite{Yang}
\begin{eqnarray}
\label{ted1}
\rho+p=\frac{1}{4\pi(4\Gamma^2-1)} \left[\mathcal{R}+K^2-K_{ij}K^{ij}-\frac{1}{2}R \right].
\end{eqnarray}
The four dimension Ricci scalar, $R$, can be decomposed as \cite{Gourgoulhon}
\begin{eqnarray}
R=\mathcal{R}+K^2+K_{ij}K^{ij}-\frac{2}{N}\mathcal{L}_m K-\frac{2}{N}D_iD^{i}N,
\end{eqnarray}
where $\mathcal{L}_m$ is the Lie derivative along $\mathbf{m}$ of any vector tangent to $\Sigma$ and $D_i$ is the Levi--Civita connection associated with the metric of the 3 dimension hypersurfaces $\Sigma$. Then Equation (\ref{ted1}) can be expressed with three dimension spacial geometrical quantities as \cite{Yang}
\begin{eqnarray}
\label{ted3}
\rho+p &=& \frac{1}{8\pi(4\Gamma^2-1)} \left[ \mathcal{R}+K^2-3K_{ij}K^{ij}+\frac{2}{N}\mathcal{L}_m K+\frac{2}{N}D_iD^{i}N \right].
\end{eqnarray}
In order to relate thermodynamics with gravity, we must suggest a hypothesis. We first review Newton's equivalence principle so as to have a better understanding. When a particle falls freely in a gravitational field, the gravity is also the inertial force. This fact leads to Newton's equivalence principle that the inertial and the gravitational mass of a particle are equal. Like the case of Newton's equivalence principle, for any perfect fluid in spacetime the energy density $\rho$ and the pressure $p$ are both the sources of gravity and thermodynamic system. So we put forward a hypothesis that the source of thermodynamic system, $\rho$ and $p$, are also the source of gravity, namely
\begin{eqnarray}
\label{hyp}
(\rho+p)_{\rm gravitational~source}=(\rho+p)_{\rm thermal~source}.
\end{eqnarray}
For radiation, this hypothesis holds \cite{Yang}. In \cite{Jacobson}, the $\delta Q$ of thermodynamic system was assumed as the energy flux and then the Einstein equation was obtained. Analogous assumptions have also been suggested in \cite{Yang,Gao2011,Wald2013,Gao2013}, though these assumptions have not been aware of by the authors. With this hypothesis (\ref{hyp}) and combining Equations (\ref{1dst}) and (\ref{ted3}), we obtain
\begin{eqnarray}
\label{key}
Ts+\mu n &=& \frac{1}{8\pi(4\Gamma^2-1)} \times \left[ \mathcal{R}+K^2-3K_{ij}K^{ij}+\frac{2}{N}\mathcal{L}_m K+\frac{2}{N}D_iD^{i}N \right].
\end{eqnarray}
The terms of the left-hand side of this equation are thermodynamical quantities, while the terms of the right-hand side of this equation are geometrical quantities of the spacetime. Equation (\ref{key}) implies that gravity can induce thermal effects, or, thermal quantities, such as entropy, temperature, and chemical potential, can induce gravitational effects. To understand the physical significance of Equation (\ref{key}) well, we consider the Newtonian approximation: $g_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}$, with $h_{\mu\nu}\ll 1$. For matter with $p\simeq0$ (the pressure of a body becomes important when its constituent particles are traveling at speeds close to that of light, which we can exclude from the Newtonian limit by hypothesis), such as dust or dark matter, we have the Poisson equation $\nabla^2\varphi=4\pi \rho$ with $\varphi=-h_{00}/2$. Taking into account Gibbs--Duhem \mbox{relation (\ref{1dst})} and the hypothesis (\ref{hyp}), we obtain
\begin{eqnarray}
\label{key1}
\nabla^2\varphi=4\pi(Ts+\mu n).
\end{eqnarray}
Equation (\ref{key1}) (or (\ref{key})) relates gravity with thermodynamics. Variations in temperature, entropy, chemical potential, and particle number will lead to a variation in the potential $\varphi$ , and vice versa. Equation (\ref{key1}) (or (\ref{key})) also indicates that gravity is related to the entropy but is not entropic force. Only for systems with constant temperature and zero chemical potential, gravitational force is entropic force. These results confirm the arguments in \cite{Kobakhidze} that experiments with ultra-cold neutrons in the gravitational field of Earth disprove the speculations on the entropic origin of gravitation. In \cite{GaoS}, it also argued that the argument for the entropic origin of gravity is problematic.
\section{Conclusions}
We have shown that if we conjecture that the source of thermodynamic system, $\rho$ and $p$, are also the source of gravity: $(\rho+p)_{\rm gravitational~source}=(\rho+p)_{\rm thermal~source}$, thermal quantities, such as entropy, temperature, and chemical potential, can induce gravitational effects, or gravity can induce thermal effects. For Newtonian approximation, the gravitational potential is related to the temperature, entropy, chemical potential, and particle number, which implies that gravity is entropic force only for systems with constant temperature and zero chemical potential. For general case, gravity is not an entropic force. Whether the results obtained here can be generalized to the case of modified gravity, such as $F(R)$ \mbox{gravity \cite{Elizalde}} and $F(\mathcal{G})$ \mbox{gravity \cite{Bamba}}, is worthy of investigation. All the analyses have been carried out without assuming a specific expression of temperature or horizon. For a static system at thermal equilibrium in general relativity, the temperature of the perfect fluid may take the form, $T\sqrt{-g_{00}}=const.$, which is called the Tolman \mbox{temperature \cite{Tolman35,Tolman36,Rovelli}}. Whether the temperature in Equation (\ref{1st}) can be taken as Tolman temperature is also worthy of further investigation. The results we obtained confirm that there is a profound connection between gravity and thermodynamics.
\acknowledgements{Acknowledgments}
This study is supported in part by National Natural Science Foundation of China (Grant Nos. 11147028 and 11273010) and Hebei Provincial Natural Science Foundation of China (Grant No. A2014201068).
\newpage
\conflictofinterests{Conflicts of Interest}
The author declares no conflict of interest.
\bibliographystyle{mdpi}
\makeatletter
\renewcommand\@biblabel[1]{#1. }
\makeatother
|
\section{Motivation}
Experimental measurements of the rare decay of a neutral pseudoscalar meson to
a lepton pair and its comparison with theoretical predictions offer an
interesting way to study low-energy (long-distance) dynamics in the Standard
Model (SM) \cite{Dorokhov:2007bd,Dorokhov:2009dg,Vasko:2011pi}. Systematical
theoretical treatment of the process dates back to 1959, when the first
prediction of the decay rate was published by Drell \cite{Drell}. While the
possible contributions of the weak sector of the SM are small enough to be
neglected, the leading order QED contribution is described by two virtual photon
exchange triangle diagram. That is why the double off-shell pion transition form
factor $F_{\pi^0\gamma^*\gamma^*}$, which is not known from the first
principles, plays essential role.
Because of this one-loop structure for the leading order, the process is very
rare and suppressed in the comparison to two photons decay
($\pi^0\rightarrow\gamma\gamma$) by a factor of 2$\left({\alpha
m_e}/{M_{\pi^0}}\right)^2$ due to the approximate helicity conservation of the
interaction and thus may be sensitive to possible effects of the physics beyond
the SM (expected branching ratio from the pure SM calculation is about
$10^{-7}$).
Recently, this decay has attracted attention of the theorists again in
connection with a new precise branching ratio measurement. The KTeV-E799-II
experiment at Fermilab \cite{Abouzaid:2006kk} has observed $\pi^0\to e^+e^-$
events (altogether 794 candidates), where $K_L\rightarrow3\pi^0$ decay was used
as a source of neutral pions. The KTeV result is
\begin{equation}
\begin{split}
\frac{\Gamma(\pi^0\to e^+e^-,\,x>0.95)}{\Gamma(\pi^0\rightarrow e^+e^-\gamma,\,x>0.232)}&=\\
=(1.685\pm0.064\pm&0.027)\times10^{-4}\,.
\end{split}
\label{eq:origexp}
\end{equation}
Here we have introduced the Dalitz variable
\begin{equation}
x
\equiv\frac{(p+q)^2}{M^2}
=\frac{(P-k)^2}{M^2}
=1-\frac{2E_k}{M}\,,
\label{eq:xEk}
\end{equation}
where $p$, $q$ and $k$ are four-momenta of electron, positron and photon,
respectively, $P=(p+q+k)$ is the four-momentum of neutral pion $\pi^0$ with a
mass $M$ and $E_k$ is the energy of the real outgoing photon in the pion CMS.
The lower bound of the Dalitz variable $x$ is used to suppress the contribution
of the Dalitz decay $\pi^0\to e^+e^-\gamma$, which naturally arises with lower
$x$.
By means of extrapolating the Dalitz branching ratio in (\ref{eq:origexp}) to
the full range of $x$, the branching ratio of the neutral pion decay into an
electron-positron pair was determined to be equal to
\begin{equation}
\begin{split}
B(\pi^0\to e^+e^-(\gamma),\,x>0.95)&=\\
=(6.44\pm0.25\pm&0.22)\times10^{-8}\,.
\end{split}
\label{eq:B}
\end{equation}
Here the first error is from data statistics alone and the second is the total
systematic error. For the matter of interest, current PDG average value
$(6.46 \pm 0.33)\times 10^{-8}$~\cite{PDG} is mainly based on this new
result.
The KTeV collaboration used the result (\ref{eq:B}) for further calculations.
They used the early calculation of Bergstr\"{o}m \cite{Bergstrom:1982wk} to
extrapolate the full radiative tail beyond $x>0.95$ and to scale the result back
up by the overall radiative corrections of 3.4\,\% to get the lowest order rate
(with the final state radiation removed) for $\pi^0\to e^+e^-$ process. The
final result is
\begin{equation}
B^{\text{no-rad}}_{\text{KTeV}}(\pi^0\to
e^+e^-)=(7.48\pm0.29\pm0.25)\times10^{-8}\,.
\end{equation}
Subsequent comparison with theoretical predictions of the SM was made in
\cite{Dorokhov:2007bd,Dorokhov:2009dg} using pion transition form factor data
from CELLO \cite{Behrend:1990sr} and CLEO \cite{Gronberg:1997fj} experiments.
Finally, it has been found, that according to SM the result should be
\begin{equation}
B^{\text{no-rad}}_{\text{SM}}(\pi^0\to e^+e^-)=(6.23\pm0.09)\times10^{-8}\,.
\end{equation}
This can be interpreted as a 3.3\,$\sigma$ discrepancy between the theory and the experiment. Of course, the discrepancy initiated further theoretical investigation of its possible sources~\cite{Dorokhov:2008qn,Kahn:2007ru}. Aside from the attempts to find the corresponding mechanism within the physics beyond the SM, also the possible revision of the SM predictions has been taken into account. Many corrections of this kind have been already made, but so far with no such a significant influence on the final result.
\section{Leading order}
According to the Lorentz symmetry the on-shell invariant matrix element of the $\pi^0\to e^+e^-$ process can be generally written in terms of just one pseudoscalar form factor
\begin{equation}
i\mathcal{M}(\pi^0\to e^+e^-)
=\overline{u}(p,m)\gamma^{5}v(q,m)P(p^2,q^2,P^2)
\end{equation}
and, as a consequence, the total decay rate is given by
\begin{equation}
\Gamma(\pi^0\to e^+e^-)
=\frac M{8\pi}\sqrt{1-\nu^2}\left|P(m^2,m^2,M^2)\right|^2\,,
\label{rate}
\end{equation}
where $m$ stands for electron mass and $\nu \equiv 2m/M$. The leading order in
the QED expansion is depicted as the left hand side of the graphical equation in
the Fig.~\ref{fig:LO}. Here the shaded blob corresponds to the off-shell pion
transition form factor $F_{\pi^0\gamma^*\gamma^*}(l^2,(P-l)^2)$ where $l$ is the
loop momentum. This form factor serves as an effective UV cut-off due to its
$1/l^2$ asymptotics governed by OPE (see e.g.~\cite{Knecht:2001xc}) and the loop
integral over $\text{d}^4l$ is therefore convergent. It is convenient to pick up
explicitly the non-analytic contribution of the two-photon intermediate state
(the imaginary part\footnote{Imaginary part of this contribution is given by
Cutkosky rules cutting the two virtual photon lines in the Fig.~\ref{fig:LO}.}
is determined uniquely up to the normalization given by the on-shell value of
$F_{\pi^0\gamma^*\gamma^*}(0,0)\equiv F_{\pi^0\gamma\gamma}$) and express the
form factor in the following way (cf. \cite{Knecht:1999gb})
\begin{equation}
\begin{split}
&P^\text{LO}(m^2,m^2,M^2)\\
&=\alpha^2mF_{\pi^0\gamma\gamma}\frac 1{\sqrt{1-\nu^2}}\left[\text{Li}_2(z)-\text{Li}_2\left(\frac 1z\right)+i\pi\log(-z)\right]\\
&+2\alpha^2mF_{\pi^0\gamma\gamma}\bigg\{\frac 32\log\left(\frac{m^2}{\mu^2}\right)-\frac 52+\chi\left(\frac{M^2}{\mu^2},\,\frac{m^2}{\mu^2}\right)\bigg\}\,.
\end{split}
\label{eq:P}
\end{equation}
Here, $\text{Li}_2$ is the dilogarithm,
\begin{equation}
z=-\frac{1-\sqrt{1-\nu^2}}{1+\sqrt{1-\nu^2}}
\end{equation
and $\mu$ represents the intrinsic scale connected with the form facto
\footnote{It means the scale at which the loop integral is effectively cut off.
The term $\frac 32\log \left( m^2/\mu^2\right)$ represents the leading
dependence of the form factor $P$ on this scale.}
$F_{\pi^0\gamma^*\gamma^*}$. The function $\chi\left(P^2/\mu^2,\,m^2/\mu^2\right)$ represents the remainder which collects the contributions of higher intermediate states and is real and analyti
\footnote{Note that the higher intermediate states, which appear when also the blob in the Fig.~\ref{fig:LO} is cut, start for $P^2\sim\mu^2$.}
for $P^2/\mu^2<1$.
\begin{figure}[t]
\centering
\include{LO_dia}
\caption{Leading order contribution in the QED expansion and its representation in terms of the leading order of the chiral perturbation theory.}
\label{fig:LO}
\end{figure}
The leading order terms in the chiral expansion of the form factor
$P^\text{LO}$ are depicted as the right hand side of the graphical equation in
Fig.~\ref{fig:LO}. The $\pi^0\gamma\gamma$ vertex in the loop graph is local and
corresponds to the leading order term of the chiral expansion of the form factor
$F_{\pi^0\gamma^*\gamma^*}$. Therefore the loop integration is no more UV finite
and a counterterm (represented by the tree graph in the Fig.~\ref{fig:LO}) is
necessary. The sum of these two terms can be written in the form (\ref{eq:P}),
where the transition form factor $F_{\pi^0\gamma\gamma}$ and the remainder
$\chi\left(P^2/\mu^2,m^2/\mu^2\right)$ are replaced by their leading orders in
the chiral expansion
\begin{equation}
F_{\pi^0\gamma\gamma}^\text{LO}
=\frac 1{4\pi^2F}\,,\;\chi^\text{LO}\left(P^2/\mu^2,m^2/\mu^2\right)
=\chi^{\text{(r)}}(\mu)\,,
\label{eq:chiLO}
\end{equation}
where $\chi^\text{(r)}(\mu)$ is the finite part of the above mentioned
counterterm renormalized at scale $\mu$. The graphical equation in the
Fig.~\ref{fig:LO} can be understood as the matching condition for
$\chi^\text{(r)}(\mu)$ at the leading order in the chiral expansion. It enables
to determine $\chi^\text{(r)}(\mu)$ once the form factor
$F_{\pi^0\gamma^*\gamma^*}$ is known. The latter can be theoretically modeled
e.g. by the lowest meson dominance (LMD) approximation to the large-$N_{C}$
spectrum of vector meson resonances yielding~\cite{Knecht:1999gb}
\begin{equation}
\chi^\text{(r)}(M_\rho)=2.2\pm0.9\,,
\label{chi22}
\end{equation}
where $M_\rho=770\,\text{MeV}$ is the mass of the $\rho$ meson. For other
alternative estimates cf. Tab.~\ref{tab:chi} and for the
complete discussion see \cite{Dorokhov:2007bd}.
\begin{table}[!ht]
\begin{center}
{\scriptsize
\begin{tabular}{c | c c c c}
\toprule
Model & \text{CLEO+OPE} & \text{QCDsr} & \text{LMD+V} & \text{N$\chi $QM}\\
\midrule
$\chi^\text{(r)}(M_\rho)$ & $2.6\pm0.3$ & $2.8\pm0.1$ & 2.5 &
$2.4\pm0.5$\\
\bottomrule
\end{tabular}}
\end{center}
\caption{Numerical values of $\chi^\text{(r)}$ in different models
according to \cite{Dorokhov:2007bd,Vasko:2011pi}. The first two columns
denoted as CLEO+OPE and QCDsr correspond to various treatments of CLEO data.
LMD+V is an improvement of the LMD ansatz and N$\chi$QM stands for the nonlocal
chiral quark model.}
\label{tab:chi}
\end{table}
Using the value (\ref{chi22}) we get for the $\pi^0\to e^+e^-$ branching
ratio numerically
\begin{equation}
B_\text{SM}^\text{LO}(\pi^0\to e^+e^-)
=(6.1\pm 0.3)\times10^{-8}\,.
\end{equation}
\section{Two-loop virtual radiative corrections}
The full two-loop virtual radiative (pure QED) corrections of order
$\mathcal{O}(\alpha^3p^2)$ were calculated
in~\cite{Vasko:2011pi}. In this section we will present a short review of the
main results.
The relevant contributions to the amplitude are shown in Fig.~\ref{fig:2loop}.
There are six two-loop diagrams. Listed sequentially, we have two vertex
corrections (a, b), electron self-energy insertion (c), box-type correction (d)
and two vacuum polarization insertions (e, f). Of course, for every such diagram
a one-loop graph with corresponding counterterm must be added to renormalize
the subdivergences. The relevant finite parts of these counterterms can be fixed
by the requirement that the parameters $m$ and $\alpha$ coincide with their
physical values. After the subdivergences are canceled, the remaining
superficial divergences has to be renormalized by another additional tree
counter-term with coupling $\xi$.
The finite part $\xi^\text{(r)}(\mu)$ of this coupling has been estimated
in~\cite{Vasko:2011pi} using its running with the renormalization scale as
\begin{equation}
\xi^\text{(r)}(M_\rho)=0\pm5.5\,.
\label{xir}
\end{equation}
\input{corrections}
Besides the UV divergences, the graph (d) in the Fig.~\ref{fig:2loop} is also
IR divergent. It is therefore necessary to consider IR-safe decay width of the
inclusive process $\pi^0\to e^+e^-(\gamma)$ with additional real photon in the
final state. In~\cite{Vasko:2011pi} the real photon bremsstrahlung has been
taken into account using the soft-photon approximation. The final result depends
on the experimental upper bound on the soft photon energy which can be expressed
in terms of the lower bound $x^\text{cut}$ on the Dalitz variable $x$ (see
(\ref{eq:xEk})). The result can be expressed in terms of the correction factor
$\delta(x^\text{cut})$ defined as
\begin{equation}
\begin{split}
&\Gamma^\text{NLO}\left(\pi^0\to e^+e^-(\gamma ),\,x>x^\text{cut}\right)\\
&\equiv\delta \left(x^\text{cut}\right)\Gamma^\text{LO}\left(\pi^0\to e^+e^-\right)\,,
\end{split}
\end{equation}
where $\Gamma^\text{LO}$ is the leading order width and $\Gamma^\text{NLO}$ is
the next-to leading $\mathcal{O}(\alpha^3p^2)$ correction. The $x^\text{cut}$
dependent overall correction $\delta(x^\text{cut})$ has various sources and to
emphasize the origin of its constituents, we will use the same symbol decorated
with appropriate
indices. For the complete QED two-loop correction $\delta^{(2)}$ including soft-photon bremsstrahlung and KTeV cut $x^\text{cut}$=$0.95$, in~\cite{Vasko:2011pi} it was obtained
\begin{equation}
\delta^{(2)}(0.95)
\equiv\delta^{\text{virt.}}+\delta_\text{soft}^\text{BS}(0.95)=(-5.8\pm0.2)\,\%\,,
\label{eq:2looprad}
\end{equation}
where only the uncertainties of $\chi^\text{(r)}$ and $\xi^\text{(r)}$ were
taken as the source of the error. This result differs significantly from the
previous approximate calculations done by Bergstr\"{o}m~\cite{Bergstrom:1982wk}
or Dorokhov et~al.~\cite{Dorokhov:2008qn}, where for
$\delta^{(2)}\left(0.95\right) $ we would get $-13.8\,\%$ and $-13.3\,\%$,
respectively.
There is a simple interrelation of this partial result of the QED radiative
corrections and the branching ratio (\ref{eq:B}) obtained by KTeV experiment
(for the details see~\cite{Vasko:2011pi}). We can write the theoretical
prediction for the branching ratio measured by KTeV as
\begin{equation}
\begin{split}
&B(\pi^0\to e^+e^-(\gamma ),\,x>0.95)
=\frac{\Gamma^\text{LO}(\pi^0\to e^+e^-)}{\Gamma (\pi^0\to\gamma\gamma)}\\
\times\,&B(\pi^0\to \gamma\gamma)\left[1+\delta^{(2)}(0.95)+\Delta^\text{BS}(0.95)+\delta^\text{D}(0.95)\right]\,,
\end{split}
\label{eq:Bth}
\end{equation}
where the only experimental input is the precise branching ratio $B(\pi^0\to\gamma\gamma)=(98.823\pm0.034)$\,\%. In the above formula,
\begin{align}
\delta^{\text{D}}(x^\text{cut})
&=\frac1{\Gamma^{\text{LO}}(\pi^0\to e^+e^-)}
{\int_{x^\text{cut}}^1}{{\text{d}}x}\left(\frac{\text{d}\Gamma^{\text{Dalitz}}}{\text{d}x}\right)_{1\gamma IR}^{\text{NLO}}\notag\\
&=\frac{1.75\times 10^{-15}}{\left[\Gamma^\text{LO}(\pi^0\to e^+e^-)/\text{MeV}\right]}
\end{align}
corresponds to the unsubtracted fraction of the Dalitz decay backgroun
\footnote{This fraction comes form the contribution of the interference term of
the NLO one-photon-irreducible ($1\gamma IR$) graph with the leading order
Dalitz amplitude. See \cite{Vasko:2011pi} and \cite{Kampf:2005tz} for more
details.}
omitted in the KTeV analysis and discussed in~\cite{Kampf:2005tz,Vasko:2011pi}.
In what follows we will concentrate on the last missing ingredient of the
formula (\ref{eq:Bth}), namely
\begin{equation}
\Delta^\text{BS}\left(x^\text{cut}\right)
\equiv\delta^\text{BS}\left(x^\text{cut}\right)-\delta_\text{soft}^\text{BS}\left(x^\text{cut}\right)\,,
\end{equation}
which is the difference between the exact bremsstrahlung and its soft photon approximation. This difference has been only roughly estimated in~\cite{Vasko:2011pi} and this estimate has been taken as a source of the error. Our aim is to calculate $\Delta^{\text{BS}}$ exactly and test the adequacy of the soft photon approximation for the cut $x^\text{cut}=0.95$ used in the KTeV analysis.
\section{Bremsstrahlung}
\label{sec:bs}
In this section, we discuss the above mentioned exact bremsstrahlung (BS), i.e.
the real radiative correction corresponding to the process $\pi^0\to
e^+e^-(\gamma)$ beyond the soft-photon approximation. As a consequence of the
gauge invariance, the invariant amplitude for the BS correction
\begin{equation}
\mathcal{M}_{(\lambda )}(p,q,k)
\equiv \varepsilon _{(\lambda )}^{* \rho }(k)\mathcal{M}_{\rho }^{\text{BS}}(p,q,k)
\end{equation}
(where $k$ and $\varepsilon _{(\lambda )}^{* \rho}(k)$ is the photon momentum
and polarization vector, respectively) has to satisfy the Ward identity
\begin{equation}
k^{\rho }\mathcal{M}_{\rho }^{\text{BS}}=0
\label{eq:BSWard}
\end{equation}
for on-shell $k$ and thus it can be generally expressed in the form~\cite{Kampf:2005tz}
\begin{equation}
\begin{split}
i&\mathcal{M}_{\rho }^{\text{BS}}(p,q,k)=\frac{i e^5}{8\pi^2 F}\times\\
&\;\;\Big\{P(x,y)\left[\left(k\cdot p\right) q_\rho-\left(k\cdot q\right) p_\rho\right]\left[\bar{u}(p,m)\gamma_5 v(q,m)\right]\\
&+A(x, y)\Big[\bar{u}(p,m)\left[\gamma_\rho\left(k\cdot p\right)-p_\rho (k\cdot\gamma)\right]\gamma_5 v(q,m)\Big]\\
&-A(x,-y)\Big[\bar{u}(p,m)\left[\gamma_\rho\left(k\cdot q\right)-q_\rho (k\cdot\gamma)\right]\gamma_5 v(q,m)\Big]\\
&+T(x,y)\left[\bar{u}(p,m)\gamma_\rho\slashed k\gamma_5 v(q,m)\right]\Big\}
\end{split}
\label{eq:BStotampl}
\end{equation}
in terms of scalar form factors $P$, $A$ and $T$. These are functions of two independent kinematic variables $(x,y)$, defined~as
\begin{equation}
\begin{gathered}
x=\frac{(p+q)^2}{M^2}\,,\;\;y=-\frac{2}{M^2}\left[\frac{k\cdot(p-q)}{1-x}\right]\\
x\in[\nu^2,1]\,,\;\;y\in\left[-\sqrt{1-\frac{\nu^2}{x}},\sqrt{1-\frac{\nu^2}{x}}\;\right]\,.
\end{gathered}
\label{eq:xybound}
\end{equation}
As mentioned above, $x$ is the Dalitz variable (i.e. a normalized square of the
total energy of $e^+e^-$ pair in their CMS) and $y$ has the meaning of a
rescaled cosine of the angle included by the directions of outgoing photon and
positron in the $e^+e^-$ CMS. The modulus squared of the amplitude has the
form~\cite{Kampf:2005tz}
\begin{equation}
\begin{split}
&\overline{\left|\mathcal{M}^\text{BS}(x,y)\right|^2}
\equiv\sum_{\text{polarizations}}\left|\mathcal{M}_{(\lambda)}(p,q,
k)\right|^2=\\
&=\frac{16\pi\alpha^5}{F^2}
\frac{M^4(1-x)^2}{8}\bigg\{
M^2\left[x(1-y^2)-\nu^2\right]\Big[xM^2\left|P\right|^2\\
&+2\nu M\operatorname{Re}\big\{P^*\left[A(x,y)+A(x,-y)\right]\big\}-4\operatorname{Re}\big\{P^*T\big\}\Big]\\
&+2M^2(x-\nu^2)(1-y)^2\left|A(x,y)\right|^2+(y\to -y)\\
&-8\nu My(1-y)\operatorname{Re}\big\{A(x,y)T^*\big\}+(y\to -y)\\
&-4\nu^2M^2y^2\operatorname{Re}\big\{A(x,y)A(x,-y)^*\big\}
+8(1-y^2)\left|T\right|^2\bigg\}
\end{split}
\label{eqMbar2}
\end{equation}
and using the variables $x$, $y$ the differential decay rate is
\begin{equation}
\text{d}\Gamma^\text{BS}(x,y)
=\frac{M}{(8\pi)^3}\overline{\left|\mathcal{M}^\text{BS}(x,y)\right|^2}(1-x)\,\text{d}x\,\text{d}y\,.
\label{eq:dGammaBS}
\end{equation}
To the amplitude $\mathcal{M}_{(\lambda)}(p,q,k)$ five Feynman diagrams
contribute (cf. Fig.~\ref{fig:BSout}). Four of them correspond to the photon
emission from the outgoing fermion lines (see
Fig.~\ref{fig:T2}---\ref{fig:T3CT}). Naively, one would expect that only these
four diagrams are necessary to consider since only they include IR divergences
which are needed to cancel the IR divergences stemming from the virtual
corrections
(see graph\ (d) in the Fig.~\ref{fig:2loop} and the corresponding one-loop
diagram
with counterterm). However, this result would not be complete.
\begin{figure}[t]
\centering
\begin{subfigure}[b]{0.4\columnwidth}
\centering
\include{dia1}
\caption{}
\label{fig:T2}
\end{subfigure}
\begin{subfigure}[b]{0.45\columnwidth}
\centering
\include{dia2}
\caption{}
\label{fig:T3}
\end{subfigure}
\begin{subfigure}[b]{0.4\columnwidth}
\centering
\include{dia1CT}
\caption{}
\label{fig:T2CT}
\end{subfigure}
\begin{subfigure}[b]{0.45\columnwidth}
\centering
\include{dia2CT}
\caption{}
\label{fig:T3CT}
\end{subfigure}
\begin{subfigure}[b]{\columnwidth}
\centering
\include{dia3}
\caption{}
\label{fig:T4}
\end{subfigure}
\caption{Bremsstrahlung Feynman diagrams for $\protect\pi^0\to e^+e^-$ process including counterterms.}
\label{fig:BSout}
\end{figure}
The reason is that the Ward identity (\ref{eq:BSWard}) would be violate
\footnote{Note that in the framework of the soft-photon approximation the sum
of these four graphs satisfies the Ward identity by itself.}.
Thus it is necessary to add the third (box) diagram (Fig.~\ref{fig:T4}, photon
emitted from the inner fermion line) to fulfill this relation.
In the graphs (\ref{fig:T2}) and (\ref{fig:T3}) the $\pi\gamma\gamma$ vertex
stems from the Wess-Zumino-Witten action \cite{Wess:1971yu,Witten:1983tw} and
the remaining vertices correspond to standard QED Feynman rules. These graphs
are UV divergent by power counting and have to be regularized. In what follows,
we use the dimensional regularization. In order to bypass the problems with
intrinsically four-dimensional objects like $\gamma_5$ and the Levi-Civita
pseudo-tensor $\varepsilon^{\mu\nu\alpha\beta}$, we use its variant known as
Dimensional Reductio
\footnote{Note however, that in general case the regularization by dimensional
reduction might spoil gauge invariance. In the case of our amplitude, we have
checked that the gauge invariance is preserved and the regularized amplitude
has the general form (\ref{eq:BStotampl}).}
(cf.~\cite{Frampton:1978ix}), which keeps the algebra of $\gamma$-matrices
four-dimensional while the loop tensor integrals are regularized dimensionally
and expressed in terms of the scalar one-loop integrals using the
Passarino-Veltman reduction \cite{Passarino:1978jh}. Within this framework we
first get rid of the Levi-Civita tensor using the four-dimensional identities,
e.g.
\begin{equation}
\begin{split}
\varepsilon^{\alpha\beta\mu\nu}\gamma_\mu\gamma_\nu&=i\gamma_5\left[\gamma^\alpha,\gamma^\beta\right]\\
\varepsilon^{\alpha\beta\mu\nu}
\gamma_\mu\gamma_\rho\gamma_\nu&=2i\gamma_5\left(g^\alpha_\rho\gamma^\beta
-g^\beta_\rho\gamma^\alpha\right)\,,
\end{split}
\end{equation}
and then contract the reduced tensor integrals with the $\gamma$-matrix structure
\footnote{According to the prescription \cite{Frampton:1978ix}, we take the metric tensors stemming from the Passarino-Veltman reduction effectively as four-dimensional.}.
The contributions of the box diagram Fig.~\ref{fig:T4} turns out to be finite
while the triangle diagrams Fig.~\ref{fig:T2} and Fig.~\ref{fig:T3} contain
subdivergences which have to be renormalized by means of the tree graphs with
counterterm corresponding to the coupling $\chi$ (see Fig.~\ref{fig:T2CT}
and~\ref{fig:T3CT}).\ Summing all the relevant contributions and using the
four-dimensional Dirac algebra, we get finally the form factors $P$, $A$, and
$T$, the explicit form of which is summarized in~\ref{app:BS}.
\begin{figure}[t]
\includegraphics[width=\columnwidth]{plot3D.eps}
\caption{3D plot of $\text{d}\Gamma_\text{diff}^\text{BS}(x,y)$ normalized to the leading order contribution of the $\protect\pi^0\to e^+e^-$ process.}
\label{fig:final}
\end{figure}
\begin{figure}[t]
\includegraphics[width=\columnwidth]{delta_x_2D.eps}
\caption{Plot of $\protect\text{d}\Gamma_\text{diff}^\text{BS}(x)=\int\text{d}\Gamma_\text{diff}^\text{BS}(x,y)\,\text{d}y$ normalized to the leading order contribution of the $\protect\pi^0\to e^+e^-$ process.}
\label{fig:x2D}
\end{figure}
The differential decay rate $\text{d}\Gamma^\text{BS}(x,y)$ (cf.~(\ref{eq:dGammaBS})) give rise to IR divergences when integrated over the phase space. The divergences originate from the soft-photon region
\begin{equation}
|\mathbf{k}|<\frac 12 M(1-x^\text{cut})\,,
\end{equation
which is defined in terms of the variables $(x,y)$ by means of the cut on the Dalitz variable $x>x^{\text{cut}}$. These divergences are exactly the same as those stemming from an analogous integral of the differential decay rate ${\text{d}}\Gamma _{\text{soft}}^{\text{BS}}(x,y)$ calculated within the soft-photon approximation. The latter is already included in the two-loop result \cite{Vasko:2011pi}, we therefore present our result for the exact BS as a difference
\begin{equation}
\text{d}\Gamma_\text{diff}^\text{BS}(x,y)
=\text{d}\Gamma^\text{BS}(x,y)-\text{d}\Gamma_\text{soft}^\text{BS}(x,y)\,,
\end{equation}
the integral of which is IR finite. The result for $\text{d}\Gamma_\text{diff}^\text{BS}(x,y)$ is shown in Fig.~\ref{fig:final} and (integrated over the allowed region of $y$ given by (\ref{eq:xybound})) in Fig.~\ref{fig:x2D}. For $\Delta ^{\text{BS}}(x^{\text{cut}})$ we get finally
\begin{equation}
\Delta^\text{BS}(x^\text{cut})
=2\int_{x^\text{cut}}^1\int_0^{\sqrt{1-\nu^2/x}}\frac{\text{d}\Gamma_\text{diff}^\text{BS}(x,y)}{\Gamma^\text{LO}\left(\pi^0\to e^+e^-\right)}\;.
\end{equation}
The dependence of $\Delta^\text{BS}(x^\text{cut})$ on $x^\text{cut}$ is shown
in Fig.~\ref{fig:xcut2D}. For $x^\text{cut}=0.95$ and for $\chi^{\text{(r)}}$
given by~(\ref{chi22}) we get numerically
\begin{equation}
\Delta ^\text{BS}(0.95)=(0.30\pm0.01)\,\%\,,
\end{equation
where the error stems from the uncertainty in $\chi^{\text{(r)}}(M_{\rho })$. In other words, using this cut of Dalitz variable in KTeV experiment, the soft-photon approximation is a very good approach to the exact result. The dependence of $\Delta^\text{BS}(0.95)$ on $\chi^{\text{(r)}}$ is shown in Fig.~\ref{fig:chi2D}.
\begin{figure}[t]
\includegraphics[width=\columnwidth]{delta_xcut_2D.eps}
\caption{The dependence of $\Delta^\text{BS}$ on the cut on the Dalitz variable.}
\label{fig:xcut2D}
\end{figure}
\begin{figure}[t]
\includegraphics[width=\columnwidth]{delta_chi_2D.eps}
\caption{The dependence of $\Delta^\text{BS}(0.95)$ on $\chi^\text{(r)}$. It is apparent, that the dependence is very slight and can be neglected in the calculation of the $\chi^\text{(r)}$.}
\label{fig:chi2D}
\end{figure}
Now we have all ingredients needed in formula~(\ref{eq:Bth}) under control
and we can thus fit the value of the coupling $\chi^\text{(r)}$
to meet the experiment with the result
\begin{equation}
\chi^\text{(r)}(M_\rho)=4.5\pm1.0\,.
\label{eq:chi}
\end{equation}
The error is dominated by the experimental uncertainty, while the
theoretical error corresponding to the estimate (\ref{xir}) is negligible.
To compare, some previously estimated values, which were considered as
relevant, are shown in Tab.~\ref{tab:chi}.
\section{Estimate of the theoretical uncertainty of~\texorpdfstring{$\chi^\text{(r)}$}{\textbackslash chi\^{}(r)}}
The above determination of $\chi^{\text{(r)}}$ represents an effective LO
value of this coupling and includes therefore implicitly higher order chiral
contributions. The corrections to the LO value of $\chi^{\text{(r)}}$
start at the NLO and stem from the two-loop graphs which correspond to a
substitution of the one-loop subgraphs (and corresponding counterterms) for the
shaded blob on the left hand side of the graphical equation depicted in Fig.~\ref{fig:LO}.
The relative size of such corrections is set by the factor $(M/4\pi
F)^{2}\sim 10^{-2}$ and can be naively treated as negligible, however it can
be significantly numerically enhanced by the large double logarithm terms
like $\log^{2}(\mu^{2}/m^{2})\sim 10^{2}$ for $\mu\sim M_{\rho}$.
\begin{figure}[!ht]
\centering
\input{plan}
\caption{One-loop diagrams of order $\alpha^2/F^3$ for $\pi^0\to e^+e^-$
process.}
\label{fig:plan}
\end{figure}
\begin{figure}[!ht]
\centering
\input{Z}
\caption{Z-factor contributions}
\label{fig:Z}
\end{figure}
Complete calculation of the NLO\ corrections is beyond the scope of the
present article. In this section, we will only restrict ourselves to the
rough estimate based on explicit calculation of the above mentioned leading
(double) logarithms, which are expected to represent a numerically
relevant part of the full NLO contribution. According to the Weinberg
consistency relation \cite{Weinberg:1978kz}, this can be achieved by means of
evaluation of infinite parts of one-loop graphs only. In what follows, we
will adapt this relation to our case.
Let us write the contribution of the above mentioned two-loop
graphs as
\begin{equation}
P^\text{NLO}
=P^\text{2-loop}+P_\text{CT}^\text{1-loop}+P_\text{CT}^\text{tree}+\left(Z^\text{1-loop}\right)^{\frac 1 2}P^\text{LO}\,,
\end{equation
where the first three terms correspond to one-particle irreducible (1PI)
contributions (including two-loop graphs, one-loop graphs with counterterms
and tree counterterm graphs) and the last term represents the
renormalization of the external pion line by means of the one-loop $Z$-factor.
The contributions of the 1PI loop graphs $P^\text{2-loop}$ can be written
schematicall
\footnote{Because we are interested only in the singular parts we ignore the
difference between $MS$, $\overline{MS}$ and $\overline{MS}_{\chi}$
subtraction schemes in what follows. Such an omission can affect only the
finite parts which are irrelevant for the leading log calculation.}
as an expansion in $\varepsilon=2-\tfrac{d}{2}$
\begin{equation}
\begin{split}
P^\text{2-loop} &=\mu^{-4\varepsilon}\left(\frac{\mu^{2}}{m^{2}}\right)^{2\varepsilon}\\
&\times\left[\frac{P_{-2}^\text{2-loop}}{\varepsilon^{2}}+\frac{P_{-1}^\text{2-loop}}{\varepsilon}+\mathcal{O}(\varepsilon^{0})\right]\,.
\end{split}
\end{equation
In the same way, for $P_\text{CT}^\text{1-loop}$ we get (see
Fig.~\ref{fig:plan})
\begin{equation}
\begin{split}
&P_\text{CT}^\text{1-loop}
=\mu^{-4\varepsilon}\left(\frac{\mu^{2}}{m^{2}}\right)^{\varepsilon}\times\\
&\Bigg[\sum_{i=7,11,13}\left(c_{i}^{W\text{(r)}}(\mu)-\frac{\eta_{i}^{W}}{32\pi^{2}\varepsilon}\right)
\left(\frac{P_{i,-1}^\text{1-loop}}{\varepsilon}+\mathcal{O}(\varepsilon^{0})\right)\\
&+\left(\chi^{\text{(r)}}(\mu)-\frac{\eta_{\chi}}{32\pi^{2}\varepsilon}\right)
\left(\frac{P_{\chi ,-1}^\text{1-loop}}{\varepsilon}+\mathcal{O}(\varepsilon^{0})\right)\Bigg]
\end{split}
\end{equation}
and the one-loop ingredients of the term $\left(Z^\text{1-loop}\right)
^{1/2}P^\text{LO}$ are then in the same way (see Fig.~\ref{fig:Z}
\begin{equation}
\begin{split}
\left(Z^\text{1-loop}\right)^\frac 12
&=\mu^{-2\varepsilon}\left[\left(\frac{\mu^{2}}{m^{2}}\right)^{\varepsilon}\Bigg(\frac{Z_{-1}^{\frac 12,\text{1-loop}}}{\varepsilon}+\mathcal{O}(\varepsilon^{0})\Bigg)\right.\\
&+\left.\beta_{4}\left( l_{4}^{\text{(r)}}(\mu)-\frac{\gamma_{4}}{32\pi^{2}\varepsilon}\right)\right]\\
P^\text{LO}
&=\mu^{-2\varepsilon}\left[\left(\frac{\mu^{2}}{m^{2}}\right)^{\varepsilon}\left(\frac{P_{-1}^\text{LO}}{\varepsilon}+\mathcal{O}(\varepsilon^{0})\right)\right.\\
&+\left.\beta_{\chi}\left(\chi^{\text{(r)}}(\mu)-\frac{\eta_{\chi}}{32\pi^{2}\varepsilon}\right)\right] .
\end{split}
\end{equation}
Here $l_{i}^{\text{(r)}}(\mu)$, $c_{i}^{W\text{(r)}}(\mu)$ and $\chi^
\text{(r)}}(\mu)$ are finite parts of the one-loop counterterms.
We use the standard notation for the two-flavour Chiral Perturbation theory
(ChPT) both in the even
\cite{Gasser:1983yg,Gasser:1984gg} and in the odd sector \cite{Bijnens:2001bb}.
The coefficient $\beta_{\chi}$ can be obtained from (\ref{eq:P}) and
(\ref{eq:chiLO})
\begin{equation}
\beta_{\chi}=\frac{1}{2}\left(\frac{\alpha}{\pi}\right)^{2}\frac{m}{F}
\end{equation}
and $\beta_4$ will be discussed below.
The Weinberg condition is based on absence of nonlocal divergences of the
form $\log\left(\mu^{2}\right) /\varepsilon $. It can be expressed as the
following constrain
\begin{equation}
\begin{split}
&0=
2P_{-2}^\text{2-loop}-\sum_{i=7,11,13}\left(\frac{\eta_{i}^{W}P_{i,-1}^\text{1-loop}}{32\pi^{2}}\right)-\frac{\eta_{\chi}P_{\chi,-1}^\text{1-loop}}{32\pi^{2}}\\
&+2Z_{-1}^{\frac 12,\text{1-loop}}P_{-1}^\text{LO}-Z_{-1}^{\frac 12,\text{1-loop}}\frac{\beta_{\chi}\eta_{\chi}}{32\pi^{2}}-\frac{\beta_{4}\gamma_{4}P_{-1}^\text{LO}}{32\pi^{2}}\,.
\end{split}
\label{eq:constr}
\end{equation}
The contribution of the leading double logs $P^{LL}$ is
\begin{equation}
\begin{split}
P^\text{LL}
&=\frac{1}{2}\log^{2}\left(\frac{\mu^{2}}{m^{2}}\right)\times\\
&\Bigg[4P_{-2}^\text{2-loop}-\sum_{i=7,11,13}\left(\frac{\eta_{i}^{W}P_{i,-1}^\text{1-loop}}{32\pi^{2}}\right)-\frac{\eta_{\chi}P_{\chi ,-1}^\text{1-loop}}{32\pi^{2}}\\
&\left. +4Z_{-1}^{\frac 12,\text{1-loop}}P_{-1}^\text{LO}-Z_{-1}^{\frac 12,\text{1-loop}}\frac{\beta_{\chi}\eta_{\chi}}{32\pi^{2}}-\frac{\beta_{4}\gamma_{4}}{32\pi^{2}}\right]\,.
\end{split}
\label{eq:log/eps}
\end{equation}
Using the constraint (\ref{eq:constr}), we get finally
\begin{equation}
\begin{split}
P^\text{LL}
&=\left(\frac{1}{8\pi}\right)^{2}\log^{2}\left(\frac{\mu^{2}}{m^{2}}\right)\Bigg[\sum_{i=7,11,13}\left(\eta_{i}^{W}P_{i,-1}^\text{1-loop}\right)\\
&+\eta_{\chi}P_{\chi ,-1}^\text{1-loop}+\beta_{\chi}\eta_{\chi}Z_{-1}^{\frac 12,\text{1-loop}}+\beta_{4}\gamma_{4}P_{-1}^\text{LO}\Bigg]\,.
\end{split}
\label{eq:PLL}
\end{equation}
Let us now discuss the ingredients of the formula (\ref{eq:PLL}).
The infinite parts of the couplings $\chi $ and $l_{4}$ are
\begin{equation}
\gamma_{4}=2\,,\;\frac{\eta_{\chi}}{32\pi^{2}}=-\frac{3}{2}\,.
\end{equation}
From the finiteness of $P^\text{LO}$, it follows
\begin{equation}
P_{-1}^\text{LO}=\frac{\beta_{\chi}\eta_{\chi}}{32\pi^{2}}=-\frac{3}{4}\left(\frac{\alpha}{\pi}\right)^{2}\frac{m}{F}\,.
\end{equation}
For the couplings $c_{i}^{W}$, the infinite parts depend on the form of the
l_{4}$ term in the chiral Lagrangian (see (see
\cite{Ananthanarayan:2002kj,Kampf:2002jh,Kampf:2009tk}) for details). For
the standard choice
\begin{equation}
\mathcal{L}_{4}^\text{std}=\frac{i l_{4}}{4}\langle u^{\mu}\chi_{\mu
-}\rangle
\end{equation
we ge
\begin{equation}
\eta_{7}^{W}=\eta_{11}^{W}=-\eta_{13}^{W}=\frac{1}{32\pi^{2}F^{2}}
\end{equation
(in this case, $\beta_{4}=0$), while for equivalent case, which
differs by terms proportional to the LO equation of motio
\begin{equation}
\begin{split}
\mathcal{L}_{4}
&=\frac{l_{4}}{8}\langle u^{\mu}u_{\mu}\rangle\langle
\chi_{+}\rangle\\
&=\frac{i l_{4}}{4}\langle u^{\mu}\chi_{\mu -}\rangle +\frac{i l_{4}}{4}\left\langle\widehat{\chi}_{-}\left(\nabla_{\mu}u^{\mu}-\frac{i}{2}\widehat{\chi}_{-}\right)\right\rangle\,,
\end{split}
\end{equation}
we get
$\beta_{4}=-(M/F)^{2}$ and
\begin{equation}
4\eta_{7}^{W}=\eta_{11}^{W}=-\eta_{13}^{W}=\frac{1}{32\pi^{2}F^{2}}\,.
\end{equation}
Because both choices have to lead to the same result, we get the following
relation
\begin{equation}
\frac{P_{7,-1}^\text{1-loop}}{(8\pi F)^{2}}=-\frac{2}{3}\beta_{4}\gamma
_{4}P_{-1}^\text{LO}=\left(\frac{\alpha}{\pi}\right)^{2}\frac{m}{F}\left(
\frac{M}{F}\right)^{2}\,.
\end{equation}
The $Z$ factor is not a physical observable therefore is both sensitive to
the field redefinition and in principle infinite. To calculate it we will use
the exponential parametrization $U=\exp\left(i\phi /F\right)$ (see e.g.
\cite{Kampf:2002jh}):
\begin{equation}
Z_{-1}^{\frac 12,\text{1-loop}}=-\frac{1}{3}\left(\frac{M}{4\pi F}\right)^{2}.
\end{equation}
The only missing ingredients are then $P_{11,-1}^\text{1-loop}$, $P_{13,-1}^\text{1-loop}$ and $P_{\chi ,-1}^\text{1-loop}$, which correspond to the
one-loop graphs depicted in the Fig.~\ref{fig:plan}. Explicitly, we get
\begin{align}
P_{\chi ,-1}^\text{1-loop}
&=\frac{2}{3}\left(\frac{\alpha}{\pi}\right)^{2}\frac{m}{F}\left(\frac{M}{4\pi
F}\right)^{2}\\
P_{11,-1}^\text{1-loop}
&=-\frac{1}{4}P_{7,-1}^\text{1-loop}\\
P_{13,-1}^\text{1-loop} &=-\left(\frac{4\pi}{3}\right)^{2}\left(\frac{\alpha}{\pi}\right)^{2}\frac{m}{F}M^{2}\left( 1-\frac{5}{2}\nu^{2}\right)\,.
\end{align}
Putting all these ingredients together, we find tha
\begin{equation}
\sum_{i=7,11}\eta_{i}^{W}P_{i,-1}^\text{1-loop}+\eta_{\chi}P_{\chi
,-1}^\text{1-loop}+\beta_{\chi}\eta_{\chi}Z_{-1}^{\frac 12,\text{1-loop}}+\beta
_{4}\gamma_{4}P_{-1}^\text{LO}=0
\end{equation}
and we get finally
\begin{equation}
\begin{split}
P^\text{LL}
&=\left(\frac{1}{8\pi}\right)^{2}\eta_{13}^{W}P_{13,-1}^\text{1-loop}\log^{2}\left(\frac{\mu^{2}}{m^{2}}\right)\\
&=\frac 1{72}\left(\frac{\alpha}{\pi}\right)^{2}\frac{m}{F}\left(\frac{M}{4\pi F}\right)^{2}\left( 1-\frac{5}{2}\nu^{2}\right)\log^{2}\left(\frac{\mu
^{2}}{m^{2}}\right) ,
\end{split}
\label{eq:PLLv}
\end{equation}
which implies the following leading log correction, which has to be
subtracted from the experimentally determined coupling (\ref{eq:chi})
\begin{equation}
\begin{split}
\Delta^\text{LL}\chi^{\text{(r)}}(\mu)
&=\beta_{\chi}^{-1}P^\text{LL}\\
&=\frac{1}{36}\left(\frac{M}{4\pi F}\right)^{2}\left( 1-\frac{5}{2}\nu^{2}\right)\log^{2}\left(\frac{\mu^{2}}{m^{2}}\right)\,.
\end{split}
\end{equation}
Numerically
\begin{equation}
\Delta^\text{LL}\chi^{\text{(r)}}(M_\rho)=0.081\,,
\end{equation}
which is well below the uncertainty of $\chi^{\rm{(r)}}$ in~(\ref{eq:chi}).
This can be taken as an indication of the robustness of our determination of
$\chi^{\rm{(r)}}$ with respect to the NLO chiral corrections.
\section{Conclusion}
In this article we have revisited the decay $\pi^0\to e^+e^-$. It has attracted
a lot of attention since its recent precise measurement by
KTeV Collaboration at Fermilab due to the discrepancy with the theoretical
predictions. Provided that the measured quantity is in agreement with the
future experiments one can attribute the existing discrepancy to the quantum
corrections, correct modeling of the double off-shell pion transition form
factor $F_{\pi^0\gamma^*\gamma^*}$ and/or possible contribution of the new
physics. Our focus here was on the first part, i.e. standard model corrections
to the leading order calculation. We have first briefly summarized recent
precise theoretical works dealing with the two-loop QED corrections. The
missing bremsstrahlung contribution to this process has been calculated. We
have
shown that the soft-photon approximation is an adequate approach in the region
of KTeV experiment.
Besides the electromagnetic corrections we have also studied possible stability
in the strong sector. It is best modeled using the higher pion-loop
contributions for example in the framework of $SU(2)$ ChPT. It is often the
case that in the two-flavour ChPT the order of these corrections can be
estimated by the size of the chiral logarithms. In fact they represent the
potential enhancement of the usual counting. We have explicitly calculated the
coefficient of the leading logarithm and due to the large suppression factor
$1/72$ (see (\ref{eq:PLLv})) it turns out to be very small. This might be an
indication of the fast convergence of the perturbation series which is
a situation similar to the chiral corrections of $\pi^0\to
\gamma^{(*)}\gamma^{(*)}$ decay (cf. \cite{Kampf:2009tk,Bijnens:2012hf}).
Using the most reliable QCD modeling of the $F_{\pi^0\gamma^*\gamma^*}$ via
the lowest-meson dominance approach \cite{Knecht:1999gb} we agree with the
estimate made in \cite{Vasko:2011pi} of $2\sigma$ discrepancy between the
theory (including all radiative corrections) and the experiment. Let us remind
that this number is significantly smaller than usually quoted difference
($3.3\sigma$), however, let us stress that this bigger number was obtained
from the rough estimates of the QED radiative corrections and it is thus an
indication of the importance of the full two-loop calculation for this process.
On the other hand, still unsatisfactory situation in the first-principle
modeling of the three-point vector-vector-pseudoscalar correlator leads to the
possibility to use the precise measurement and the full radiative calculation
of this process to set the hadronic form factor, represented for this process
by the constant $\chi$. The obtained value $\chi^\text{(r)}(M_\rho)=4.5\pm1.0$
(see (\ref{eq:chi})) is slightly different from the usual estimations, however,
represents the model independent prediction for this quantity, based on the
KTeV experiment. It can be further used e.g. in the hadronic light-by-light
contribution of the muon $g-2$ (see e.g.
\cite{RamseyMusolf:2002cy,Miller:2012opa} for details).
\section*{Acknowledgment} This work is supported
by Charles University in Prague, project PRVOUK P45, and by Ministry of
Education of the Czech Republic, grant LG 13031. T.H. was supported by the
grants SVV 260097/2014 and GAUK 700214.
|
\section{Introduction}
If you roll a fair die many times and compute the average number of spots showing, the result is likely to be close to 3.5, and the odds that the average is far from the expected value decreases roughly as the area under the familiar bell-shaped curve. Something similar happens if the measurement is continuous rather than discrete, such as when you repeatedly toss a needle on the ground and measure the angle it makes with respect to a fixed reference line. Even if the die is not a fair die or the geometry of the needle and the ground makes some angles more likely than others, the distribution of the average still approaches the area under a bell-shaped curve centered on the expected value. The width of the bell depends on both the variance of the random measurement and the number of times it is performed. Made precise, this amounts to a statement of the Central Limit Theorem.
The theorem was proved by De Moivre in the eighteenth century in the special case where the repeated experiment is a toss of a fair coin. It was generalized by Laplace in the early nineteenth century, and later by Cauchy and others. However, the the first rigorous proof of the Central Limit Theorem at level of generality hinted at above was given by Lyapunov in 1901.
Even in more rudimentary formulations, the Central Limit Theorem says, in a sense, that the behavior of random or chaotic events becomes highly predictable when averaged over the long term. In 1889, Sir Francis Galton waxed poetic about this conclusion:
\begin{quote}
I know of scarcely anything so apt to impress the imagination as the wonderful form of cosmic order expressed by the ``Law of Frequency of Error.'' The law would have been personified by the Greeks and deified, if they had known of it. It reigns with serenity and in complete self-effacement, amidst the wildest confusion. The huger the mob, and the greater the apparent anarchy, the more perfect is its sway. It is the supreme law of Unreason. Whenever a large sample of chaotic elements are taken in hand and marshaled in the order of their magnitude, an unsuspected and most beautiful form of regularity proves to have been latent all along. \cite[page 66]{galton:89}
\end{quote}
The result lies at the heart of modern probability theory. Many generalizations and variations have been studied; for example, relaxing the requirement that the repeated measurements are independent of one another and identically distributed, or providing additional information on the rate of convergence.
In this note, we report on a formalization of the Central Limit Theorem in the Isabelle proof assistant. We feel that this formalization is interesting and valuable for a number of reasons. Not only is the Central Limit Theorem fundamental to contemporary probability theory and the study of stochastic processes, but so is almost all of the machinery developed to prove it, ranging from ordinary calculus to the properties of real distributions and characteristic functions. There is a pragmatic need to have statistical claims made in engineering, risk analysis, and financial computation subject to formal verification, and the library we have built provides infrastructure for such practical efforts.
The formalization is also a good test for Isabelle's libraries, proof language, and automated reasoning tools. As will become clear below, the proof draws on a very broad base of facts from analysis, topology, measure theory, and probability theory, providing a useful evaluation of the robustness and completeness of the supporting libraries. Moreover, the concepts build on one another. For example, a measure is a function from a class of sets to the reals, and reasoning about convergence of measures involves reasoning about sequences of such functions. The operation of forming the characteristic function is a functional taking a measure to a function from the reals to the complex numbers, and the convergence of such functionals is used to deduce convergence of measures. Thus the conceptual underpinnings are as deep as they are broad, and working with the notions exercises Isabelle's mechanisms for handling abstract mathematical notions.
Although the Central Limit Theorem, as formulated in the next section, has been completely verified, the proofs and supporting libraries are still under development. The proofs can be found online at \url{https://github.com/avigad/isabelle}. This report provides only a brief overview, and we will provide more detail in a subsequent version.
\section{The measure-theoretic formulation}
For our formalization, we followed Billingsley's oft-used textbook, \emph{Probability and Measure} \cite{billingsley:95}, which provides an excellent introduction to these subjects. Here we briefly review the core concepts and give a precise statement of the Central Limit Theorem.
A \emph{measure space} $(\Omega, \mdl F)$ consists of a set $\Omega$ and a \emph{$\sigma$-algebra $\mdl F$ of subsets of $\Omega$}, that is, a collection of subsets of $\Omega$ containing the empty set and closed under complements and countable unions. Think of $\Omega$ as the set of possible states of affairs, or possible outcomes of an action or experiment, and each element $E$ of $\mdl F$ as representing the set of states or outcomes in which some \emph{event} occurs --- for example, that a card drawn is a face card, or that Spain wins the World Cup. A \emph{probability measure} $\mu$ on this space is a function that assigns a value $\mu(E)$ in $[0, 1]$ to each event $E$, subject to the following two conditions:
\begin{enumerate}
\item $\mu(\emptyset) = 0$, and
\item $\mu$ is countably additive: if $(E_i)$ is any sequence of disjoint events in $\mdl F$, $\mu(\bigcup_i E_i) = \sum_i \mu(E_i)$.
\end{enumerate}
Intuitively, $\mu(E)$ is the ``probability'' that $E$ occurs.
The collection $\mdl B$ of \emph{Borel subsets} of the real numbers is the smallest $\sigma$-algebra containing all intervals $(a, b)$. A \emph{random variable} $X$ on the measure space $(\Omega, \mdl F)$ is a measurable function from $(\Omega, \mdl F)$ to $(\mathbb{R}, \mdl B)$. Saying $X$ is measurable means that for every Borel subset $B$ of the real numbers, the set $\{ \omega \in \Omega \; | \; X(\omega) \in B \}$ is in $\mdl F$. Think of $X$ as some real-valued measurement that one can perform on the outcome of the experiment, in which case, the measurability of $X$ means that if we are given any probability measure $\mu$ on $(\Omega, \mdl F)$, then for any Borel set $B$ it makes sense to talk about ``the probability that $X$ is in $B$.'' In fact, if $X$ is a random variable, then any measure $\mu$ on $(\Omega, \mdl F)$ gives rise to a measure $\nu$ on $(\mathbb{R}, \mdl B)$, defined by $\nu(B) = \mu ( \{ \omega \in \Omega \; | \; X (\omega) \in B \})$. A measure on $(\mathbb{R}, \mdl B)$ is called a \emph{real distribution}, or, more simply, a \emph{distribution}, and the measure $\nu$ just described is called \emph{the distribution of $X$}.
If $X$ is a random variable, the \emph{mean} or \emph{expected value} of $X$ with respect to a probability measure $\mu$ is $\int X d\mu$, the integral of $X$ with respect to $\mu$. If $m$ is the mean, the \emph{variance} of $X$ is $\int (X - m)^2 d\mu$, a measure of how far, on average, we should expect $X$ to be from its average value.
Note that passing from $\mu$ and $X$ to its distribution $\nu$ means that instead of worrying about the probability that some abstract event occurs, we focus more concretely on the probability that some measurement on the outcome lands in some set of real numbers. In fact, many theorems of probability theory do not really depend on the abstract space $(\Omega, \mdl F)$ on which $X$ is defined, but, rather, the associated distribution on the real numbers. Nonetheless, it is often more intuitive and convenient to think of the real distribution as being the distribution of a random variable (and, indeed, any real distribution can be represented that way).
One way to define a real distribution is in terms of a \emph{density}. For example, in the case where $\Omega = \{1, 2, 3, 4, 5, 6\}$, we can specify a probability on all the subsets of $\Omega$ by specifying the probability of each of the events $\{1\}, \{2\}, \ldots, \{6\}$. More generally, we can specify a distribution $\mu$ on $\mathbb{R}$ by specifying a function $f$ such that for every interval $(a, b)$, $\mu((a, b)) = \int_a^b f x \; \mathit{dx}$. The measure $\mu$ is then said to be the real distribution with density $f$. In particular, the \emph{normal distribution} with mean $m$ and variance $\sigma^2$ is defined to be the real distribution with density function
\[
f(x) = \frac{1}{\sigma \sqrt{2 \pi}} e^\frac{-(x - m)^2}{2 \sigma^2},
\]
a ``bell shaped curve'' centered at $m$. When $m = 0$ and $\sigma = 1$, this is called the \emph{standard normal distribution}.
Let $X_0, X_1, X_2, \ldots$ be any sequence of independent random variables, each with the same distribution $\mu$, mean $m$, and variance $\sigma^2$. Here ``independent'' means that the random variables $X_0, X_1, \ldots$ are all defined on the same measure space $(\Omega, \mdl F)$, but they represent independent measurements, in the sense that for any finite sequence of events $B_1, B_2, \ldots, B_k$ and any sequence of distinct indices $i_1, i_2, \ldots, i_k$, the probability that $X_{i_j}$ is in $B_j$ for each $j$ is just the product of the individual probabilities that $X_{i_j}$ is in $B_j$. For each $n$, let $S_n = \sum_{i < n} X_i$. Notice that each $S_n$ is really a measurable function on $(\Omega, \mdl F)$, which is to say, it is a random variable; and so it is natural to ask how its values are distributed. We can shift the expected value of $S_n$ to $0$ by subtracting $n m$, and scale the variance to $1$ by dividing by $\sqrt{ n \sigma^2}$. The Central Limit Theorem says that the corresponding quantity,
\[
\frac{S_n - nm}{\sqrt{n \sigma^2}},
\]
approaches the standard normal distribution as $n$ approaches infinity.
All that remains to do is to make sense of the assertion that a sequence of distributions $\mu_0, \mu_1, \mu_2, \ldots$ ``approaches'' a distribution, $\mu$. For distributions that are defined in terms of densities, the intuition is that over time the graph of the density should look more and more like the graph of the density of the limit. For example, if you flip a coin a number of times and graph all the possible values of the average number of ones, the discrete points plotted over the possibilities $0, 1/n, 2/n, 3/n, \ldots, 1$ start to look like a bell-shaped curve centered on $1 / 2$. The notion of \emph{weak convergence} makes the notion of ``starts to look like'' precise.
If $\mu$ is any real distribution, then the function $F_\mu(x) = \mu((-\infty, x])$ is called the \emph{cumulative distribution function} of $\mu$. In words, for every $x$, $F_\mu(x)$ returns the likelihood that a real number chosen randomly according to the distribution is at most $x$. Clearly $F_\mu(x)$ is nondecreasing, and it is not hard to show that $F_\mu$ is right continuous, approaches $0$ as $x$ approaches $-\infty$, and approaches $1$ as $x$ approaches $\infty$. Conversely, one can show that any such function is the cumulative distribution function of a unique measure. Thus there is a one-to-one correspondence between functions $F$ satisfying the properties above and real distributions.
The notion of weak convergence can be defined in terms of the cumulative distribution function:
\begin{definition}
Let $(\mu_n)$ be a sequence of real distributions, and let $\mu$ be a real distribution. Then \emph{$\mu_n$ converge weakly $\mu$}, written $\mu_n \Rightarrow \mu$, if $F_{\mu_n}(x)$ approaches $F_\mu(x)$ at each point $x$ where $F_\mu$ is continuous.
\end{definition}
To understand why we need to exclude points of continuity of $F_\mu$, for each $n$, consider the probability measure $\mu_n$ that puts all its ``weight'' on $1 / n$, which is to say, for any Borel set $B$, $\mu(B) = 1$ if and only if $B$ contains $1 / n$. Then $F_{\mu_n}$ is the function that jumps from $0$ to $1$ at $1 / n$. Intuitively, it makes sense to say that $\mu_n$ approaches the real distribution $\mu$ that puts all its weight at $0$. But for every $n$, $F_{\mu_n}(0) = 0$, while $F_\mu(0) = 1$, which explains why want to exclude the point $0$ from consideration. Notice that since $F_\mu$ is a monotone function, it can have at most countably many points of discontinuity, so we are excluding only countably many points.
That the notion of weak convergence is robust is supported by the fact that there are a number of equivalent characterizations. The following theorem is sometimes known as the \emph{Portmanteau Theorem}:
\begin{theorem}
The following are equivalent:
\begin{itemize}
\item $\mu_n \Rightarrow \mu$
\item $\int f \; d\mu_n$ approaches $\int f \; d\mu$ for every bounded, continuous function $f$
\item If $A$ is any Borel set, $\partial A$ denotes the topological boundary of $A$, and $\mu(\partial A) = 0$, then $\mu_n(A)$ approaches $\mu_n(A)$.
\end{itemize}
\end{theorem}
The Central Limit Theorem can now be stated precisely as follows:
\begin{theorem}
\label{theorem:clt}
Let $X_0, X_1, X_2, \ldots$ be a sequence of independent random variables with mean $0$, variance $\sigma^2$, and common distribution $\mu$. Let $S_n = (X_0 + X_1 + \ldots + X_{n-1})$. Then the distribution of $S_n / \sqrt{n \sigma^2}$ converges weakly to the standard normal distribution.
\end{theorem}
The restriction to mean $0$ does not result in any loss of generality, since if each $X_i$ has mean $m$, we can apply Theorem~\ref{theorem:clt} to the sequence $(X_i - m)$ to obtain the more general statement above. Our formulation of Theorem~\ref{theorem:clt} in Isabelle is depicted as Figure~\ref{fig:clt}.
\begin{figure}
\begin{isabellebody}
\isacommand{theorem}\isamarkupfalse%
\ {\isacharparenleft}\isakeyword{in}\ prob{\isacharunderscore}space{\isacharparenright}\ central{\isacharunderscore}limit{\isacharunderscore}theorem{\isacharcolon}\isanewline
\ \ \isakeyword{fixes}\ \isanewline
\ \ \ \ X\ {\isacharcolon}{\isacharcolon}\ {\isachardoublequoteopen}nat\ {\isasymRightarrow}\ {\isacharprime}a\ {\isasymRightarrow}\ real{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ {\isasymmu}\ {\isacharcolon}{\isacharcolon}\ {\isachardoublequoteopen}real\ measure{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ {\isasymsigma}\ {\isacharcolon}{\isacharcolon}\ real\ \isakeyword{and}\isanewline
\ \ \ \ S\ {\isacharcolon}{\isacharcolon}\ {\isachardoublequoteopen}nat\ {\isasymRightarrow}\ {\isacharprime}a\ {\isasymRightarrow}\ real{\isachardoublequoteclose}\isanewline
\ \ \isakeyword{assumes}\isanewline
\ \ \ \ X{\isacharunderscore}indep{\isacharcolon}\ {\isachardoublequoteopen}indep{\isacharunderscore}vars\ {\isacharparenleft}{\isasymlambda}i{\isachardot}\ borel{\isacharparenright}\ X\ UNIV{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ X{\isacharunderscore}integrable{\isacharcolon}\ {\isachardoublequoteopen}{\isasymAnd}n{\isachardot}\ integrable\ M\ {\isacharparenleft}X\ n{\isacharparenright}{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ X{\isacharunderscore}mean{\isacharunderscore}{\isadigit{0}}{\isacharcolon}\ {\isachardoublequoteopen}{\isasymAnd}n{\isachardot}\ expectation\ {\isacharparenleft}X\ n{\isacharparenright}\ {\isacharequal}\ {\isadigit{0}}{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ {\isasymsigma}{\isacharunderscore}pos{\isacharcolon}\ {\isachardoublequoteopen}{\isasymsigma}\ {\isachargreater}\ {\isadigit{0}}{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ X{\isacharunderscore}square{\isacharunderscore}integrable{\isacharcolon}\ {\isachardoublequoteopen}{\isasymAnd}n{\isachardot}\ integrable\ M\ {\isacharparenleft}{\isasymlambda}x{\isachardot}\ {\isacharparenleft}X\ n\ x{\isacharparenright}\isactrlsup {\isadigit{2}}{\isacharparenright}{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ X{\isacharunderscore}variance{\isacharcolon}\ {\isachardoublequoteopen}{\isasymAnd}n{\isachardot}\ variance\ {\isacharparenleft}X\ n{\isacharparenright}\ {\isacharequal}\ {\isasymsigma}\isactrlsup {\isadigit{2}}{\isachardoublequoteclose}\ \isakeyword{and}\isanewline
\ \ \ \ X{\isacharunderscore}distrib{\isacharcolon}\ {\isachardoublequoteopen}{\isasymAnd}n{\isachardot}\ distr\ M\ borel\ {\isacharparenleft}X\ n{\isacharparenright}\ {\isacharequal}\ {\isasymmu}{\isachardoublequoteclose}\isanewline
\ \ \isakeyword{defines}\isanewline
\ \ \ \ {\isachardoublequoteopen}S\ n\ {\isasymequiv}\ {\isasymlambda}x{\isachardot}\ {\isasymSum}i{\isacharless}n{\isachardot}\ X\ i\ x{\isachardoublequoteclose}\isanewline
\ \ \isakeyword{shows}\isanewline
\ \ \ \ {\isachardoublequoteopen}weak{\isacharunderscore}conv{\isacharunderscore}m\ {\isacharparenleft}{\isasymlambda}n{\isachardot}\ distr\ M\ borel\ {\isacharparenleft}{\isasymlambda}x{\isachardot}\ S\ n\ x\ {\isacharslash}\ sqrt\ {\isacharparenleft}n\ {\isacharasterisk}\ {\isasymsigma}\isactrlsup {\isadigit{2}}{\isacharparenright}{\isacharparenright}{\isacharparenright}\ \isanewline
\ \ \ \ \ \ \ \ {\isacharparenleft}density\ lborel\ standard{\isacharunderscore}normal{\isacharunderscore}density{\isacharparenright}{\isachardoublequoteclose}
\end{isabellebody}
\caption{The Central Limit Theorem}
\label{fig:clt}
\end{figure}
\section{Supporting libraries}
In this section, we discuss some of the libraries of theorems in Isabelle that supported our formalization, and some of the ways we found it useful or necessary to extend those libraries.
To start with, Isabelle's libraries include a development of the notion of a sigma-additive measure on a space, Lebesgue integration with respect to any such measure, and standard properties of integration, including the monotone convergence theorem, the dominated convergence theorem, and Fubini's theorem. For non-finite measures, such as Lebesgue measure on the real line, measures and integrals take values in the \emph{extended reals}, which include $\pm \infty$. Probability measures, however, are instance of finite measures, which always take values in the reals. The measure theory library also includes the definition of an independent family of sets and an independent family of random variables, and some of their basic properties. These were crucial to our formalization. This library is described in \cite{hoelzl2012thesis,hoelzl2011measure}, and served our purposes well, providing, in particular, automated methods for establishing tedious measurability claims. The formalization helped us round out the library and fill some minor omissions, adding to its robustness.
When $\mu$ is a measure, we found it useful to introduce notation for the integral $\int_A f \; d\mu$ over a set $A$. Using the characteristic function of $A$, this is interpreted as $\int f \cdot 1_A \; d\mu$. We also found it useful to restate common facts from the integration library (addition formulas, etc.) ``relativized'' to a set. In hindsight, it would have been smoother to take integration over a set to be a basic notion, and then take $\int f \; d\mu$ to be notation for integration over the entire type in question.
We also drew on Isabelle's library of topological notions, including limits and continuity. Isabelle also has powerful library for dealing with limits, described in \cite{hoelzl2013typeclasses}, which supports reasoning about limits of sequences and limits of functions. In the latter case, the relevant concepts are general enough to allow both the arguments and the result to approach a particular real number, or infinity, or negative infinity. Moreover, the domain and range can be restricted, for example, allowing us to reason about limits and continuity from the left or right. These notions are central to our formalization.
Isabelle library includes a good development of notions from point-set topology. We were primarily concerned topological properties of the real numbers. For example, recall that one of the equivalence characterizations of the weak convergence of measures asserts that the sequence converges on any measurable set whose boundary has measure $0$. Facts about cardinality also came into play. For example, we had to prove that a monotone increasing function on the reals has only countably many points of discontinuity, and hence, since any nonempty interval is uncountable, one can find a point in any such interval at which the function is continuous.
Isabelle's analysis library includes a theory of differentiation, and standard properties of sine, cosine, and exponentiation, and their Taylor series expansions. There is also a development of multivariate analysis, based on John Harrison's development of the notions in HOL light. This includes a formalization of the Fr\'echet derivative for arbitrary real normed vector spaces, and the gauge integral on any Euclidean space. We only needed the univariate versions. Note that this means that there are two distinct versions of integration, since there is also the measure-theoretic notion, defined for functions from any measure space to the reals. There is some overlap, however; for example, the library shows that the two integrals, taken over a finite integral, are equal for Riemann integrable functions on that interval.
For our formalization, we had to make extensive use of ordinary calculus, for example, calculating $\int_0^\infty (\sin x) / x \; dx$ and $\int_0^\infty x^k e^{-x^2} \; dx$ for every $k$. These are tricky because the limits of integration make them ``improper'' integrals. Textbook presentations of these calculations rely explicitly or implicitly on various appeals to limits, and one has to take care to ensure that expressions in questions are integrable. Both calculations involve changes of variable (the ``substitution theorem'' from calculus) and integration by parts, and the first calculation relies on Fubini's theorem. To support the calculations, we defined the notion of integration over an interval, $\int_a^b f(x) \; dx$, with two interesting twists: $a$ and $b$ are extended reals, which is to say, they can be $\pm \infty$, and also $b$ is allowed to be less than $a$, in which case $\int_a^b f(x) \; dx = - \int_b^a f(x) \; dx$. We developed versions of the fundamental theorem of calculus, the substitution lemma, and so on, compatible with this notion. In particular, our library supports arguments where one calculates an integral $\int_a^b f(x) \; dx$, and then allows $a$ and $b$ to approach the endpoints of the interval of interest (for example, letting them go off to $\pm \infty$, respectively). (For our purposes, it was sufficient to have the substitution lemma for integrals of continuous functions, but Manuel Eberl has recently generalized the result to arbitrary measurable functions.)
We also extended the library for measure theory and probability to include many of the facts mentioned in the last section. Establishing the correspondence between real distributions and cumulative distribution functions
required a compactness argument and the Carath\'eodory extension theorem. Proving the harder implications of the Portmanteau Theorem required proving \emph{Skorohod's theorem}, a generally useful fact that shows that under general conditions it is possible to represent a sequence of measures on a common space. This required a long and delicate argument.
Finally, as will become clear in the next section, we also need notions of integration and differentiation for functions from $\mathbb{R}$ to $\mathbb{C}$, the complex numbers. This is a starting point for the development of complex analysis (e.g.~as carried out by Harrison \cite{harrison:07}), but it is much simpler: the integral of a function $f : \mathbb{R} \to \mathbb{C}$ is simply defined in terms of the integral of its real and imaginary parts, and similarly for the derivative. Since the complex numbers can be viewed as a vector space over the reals, we could show that the natural definition of the complex derivative of a function $f : \mathbb{R} \to \mathbb{C}$ coincides with derivative of $f$ considered as function from $\mathbb{R}$ to a real-valued vector space. For the integral, however, we had to define the new notion and ``import'' basic properties of real-valued integrals to the complex version (in addition, of course, to establishing properties that are specific to the complex integral). In fact, the three notion of integral we have considered here (the multivariate version, the version for functions from an arbitrary measure space to the reals, and the version for functions from an arbitrary measure space to the complex numbers) admit a common generalization, which would unify the library substantially.
\section{Overview of the proof}
Contemporary proofs of the Central Limit Theorem rely on the use of \emph{characteristic functions}, a powerful method that dates back to Laplace. If $\mu$ is a real-valued distribution, its characteristic function $\varphi(t)$ is defined by
\[
\varphi(t) = \int_{-\infty}^{\infty} e^{itx} \mu(dx).
\]
In words, $\varphi(t)$ is the integral of the function $f(x) = e^{itx}$ over the whole real line, with respect to $\mu$. Notice that for each $t \neq 0$, the function $e^{itx}$ is periodic with period $2 \pi / t$. It might be helpful to think of $e^{itx}$, as a function of $x$, as like a sine or cosine in $x$ whose period depends on $t$. (Indeed, $e^{itx}= \cos (t x) + i \sin (t x)$.) Notice that $\varphi(0) = 1$, the measure of the entire real line. The characteristic function of a real distribution $\mu$ is a Fourier transform of the measure $\mu$, and when $t \neq 0$, $\varphi(t)$ ``detects'' periodicity in the way that the real distribution $\mu$ distributes its ``weight'' over different parts of the real line.
The \emph{Levy Uniqueness Theorem} asserts that if $\mu_1$ and $\mu_2$ have the same characteristic function, then $\mu_1 = \mu_2$. In other words, a measure $\mu$ can be ``reconstructed'' from its characteristic function, and so the characteristic function of a measure determines the measure uniquely. To prove the uniqueness theorem, it is enough to show that $\mu_1((a,b]) = \mu_2((a,b])$ for real numbers $a$ and $b$ with the property that $\mu_1(\{a\}) = \mu_2(\{a\}) = \mu_1(\{b\}) = \mu_2(\{b\}) = 0$. The proof involves approximating the indicator function of $(a,b]$ in terms of the complex exponential (this is where the function $(\sin x) / x$ comes into play), unfolding the definition of the characteristic function, and doing an explicit calculation with integrals and limits.
Let $(\mu_n)$ be a sequence of distributions, where each $\mu_n$ has characteristic function $\varphi_n$, and let $\mu$ be a distribution with characteristic function $\varphi$. The \emph{Levy Continuity Theorem} states that $\varphi_n$ converges to $\varphi$ weakly if and only if $\varphi_n(t)$ converges to $\varphi(t)$ for every $t$. Proving the ``only if'' direction is easy, but the other direction is a lot harder. It relies on a result known as \emph{Helly's Selection Theorem}, a form of compactness for the space of distributions.
Remember that the CLT asserts that if $(X_n)$ is a sequence of random variable satisfying certain hypotheses, and, for each $n$, $\mu_n$ is a certain distribution defined in terms of $X_1, \ldots, X_n$, then $\mu_n$ converges weakly to the standard normal distribution. The considerations of the previous paragraph provide a simple strategy for proving the theorem: writing the characteristic function of $\mu_n$ as $\varphi_n$ for each $n$, we need only show that $\varphi_n$ approaches the characteristic function of the standard normal distribution pointwise.
To implement this strategy, one needs to know that the characteristic function of the standard normal distribution is $\varphi(t) = e^{-t^2/2}$. Establishing this fact took more work than we thought it would. Many textbook proofs of this invoke facts from complex analysis that were unavailable to us. Billingsley \cite[page 344]{billingsley:95} sketches an elementary proof, which required calculating the moments and absolute moments of the standard normal distribution. (This is where the calculations of $\int_{-\infty}^\infty x^k e^{-x^2 / t} \; dx$, mentioned in the last section, were needed. A prior calculation by Sudeep Kanav covered the cases $k = 0, 1$, which provide the base cases for an inductive proof.) Filling in the details involved carrying out careful computations with integrals and power series approximations to $e^x$.
The other component to implementing the strategy involves computing the characteristic functions of the distributions $\mu_n$, which are defined in terms of finite sums of the random variables $X_0, X_1, \ldots$. This is where the utility of characteristic functions becomes apparent. If $X$ is a random variable, define the characteristic function of $X$ to be the characteristic function of its distribution. A straightforward calculation involving properties of independent random variables shows that if $X_1, \ldots, X_n$ are independent random variables then the characteristic function of $X_1 + \ldots + X_n$ is the product of the characteristic functions of each $X_i$. This facilitates the computation of the characteristic functions of the distributions $\mu_n$.
Once all these components are were in place, putting the pieces together was not hard. Given the continuity theorem, the characteristic function of the standard normal distribution, the result on the characteristic functions of sums of random variables, and a power series approximation to the complex exponential function, the proof of the Central Limit Theorem is quite short. In our formalization, it was only about 120 lines long.
It is a mistake to judge the complexity of a formalization by the number of lines. For one thing, different styles of writing proofs can compress or expand the number of lines dramatically. Moreover, length is not a good proxy for difficulty: some proofs are short but hard (for example, because they involve combining heterogeneous facts in tricky ways) and others are long and easy. For what it is worth, however, the files in our repository contain about 13,000 lines.
\bibliographystyle{plain}
|
\section{Introduction}
The cosmic microwave background (CMB) is a powerful tool to probe
the early universe and the cosmological evolution that followed. Temperature
fluctuations have been very well measured by independent experiments
\cite{Planck_params2013, 2012arXiv1212.5225B}, including the first cosmological
data released by the Planck collaboration \cite{Ade:2013ktc}. The data are in agreement
with the $\Lambda$CDM model, although still compatible with
scenarios beyond the standard framework (see for example \cite{Pettorino_2013}). For this reason, it is important to study the potential of new observables than can help to discriminate different models and constrain them further. An obvious possibility are CMB B-modes. Thomson scattering in the presence of primordial fluctuations affects
not only the temperature of the CMB but also its polarization. B-modes generated by gravitational
lensing of the CMB by large scale structure (LSS) have been observed by
two independent teams \cite{Hanson:2013hsb,Ade:2013gez}. In addition,
the BICEP2 collaboration has recently claimed the detection of a B-mode signal
in the CMB (around $\ell\sim80$), interpreted by the
team as the imprint of primordial gravitational waves from inflation
\cite{Ade:2014xna}. Whether this is indeed the case will depend on the analysis
of independent probes, the cross-correlation of data at different
frequencies and a careful check of foregrounds of polarized dust emission
\cite{2012arXiv1212.5225B,Liu:2014mpa,2014arXiv1405.0874P}.
Cosmological models in which gravity is modified with respect
to Einstein's General Relativity (GR) affect the CMB spectra in several
ways \cite{amendola_etal_2012}, for example through the Late Integrated
Sachs-Wolfe effect \cite{Giannantonio:2009zz} and CMB lensing of
the temperature spectrum \cite{acquaviva_baccigalupi_2006,2004PhRvD..70b3515A}.
In this paper we study how modified gravity (MG) affects B-mode spectra and
identify two main observable effects.
The first effect concerns the lensing
contribution to B-modes and is generated by the anisotropic stress
that characterizes modified gravity theories. Gravitational lensing of the CMB by the large-scale structure (LSS)
affects temperature anisotropies and its ``electric'' (E) and ``magnetic''
(B) polarized components \cite{Lewis_Challinor_2006}. The presence of anisotropic stress in MG generically affects
the lensing potential and changes the TT, EE and BB spectra (and the cross-spectra) with respect to those predicted by GR and $\Lambda$CDM.
The second effect is related to the speed $c_{T}$ of gravitational waves, which can
change the peak position of primordial B-modes because $c_T$ determines their epoch of horizon crossing. Once foreground emission
will be under control, these two effects will give a novel handle
for using future B-mode measurements as a tool to test MG.
There are important difficulties that arise in trying to test
these two effects. First of all, most MG models affect at the same time
the lensing potential and the matter Poisson equation, thereby a change
in the polarization spectra will typically appear associated to a
modification of the growth rate of scalar perturbations. Since the
latter begins to be considerably constrained by the current data (see
e.g. \cite{Macaulay:2013swa}) and the temperature spectrum of scalar
perturbations in the CMB is very well determined \cite{Planck_params},
it could be naively expected that a substantial modification of the
BB-spectrum would also imply a large imprint in the growth
of structures or the TT spectrum. In addition, there is a large
variety of MG models (see for ex. \cite{Amendola2010,Clifton:2011jh,Yoo:2012ug})
and adapting a CMB Boltzmann code for a significant fraction of them
is a daunting task. Conversely, restricting the analysis to a narrow
subset of cases would reduce the appeal of a study of B-modes in MG.
In this paper we propose to bypass these problems by adopting a radical
strategy. We will focus on MG models with a single extra scalar degree
of freedom in the linear quasi-static regime, i.e. taking the large
$k$ limit and neglecting the time derivatives of the potentials.
This reduces considerably the complexity of vast classes of models.
Concretely, this can be applied to the entire class of Horndeski Lagrangians
\cite{Horndeski:1974,Deffayet:2011gz}, which comprises most viable modified gravity models based on a scalar field, and bimetric gravity models
\cite{Hassan:2011zd}. Furthermore, we assume
that the modification of the Poisson equation is negligible, effectively
selecting those models whose effect on the CMB is essentially due
to lensing and the propagation of gravitational waves. We also assume
that the background behaviour is the one of $\Lambda$CDM. In other words, we assume that MG only affects those observables related to gravitational lensing and gravitational
waves. This choice allows us to test the effects of the anisotropic stress
and the sound speed of gravitational waves with a prescription that can be easily implemented
in Boltzmann codes.
MG effects are in general time dependent. We can therefore test their
impact at various epochs during the history of the universe (e.g.
at decoupling time or at later times). Lensing effects are mostly
generated at a redshift of order unity, while the speed of gravitational
waves affects the CMB at decoupling time and, to a minor extent, at or before
reionization. In the following we will thus treat separately the following
cases: MG effects present at all times, only at decoupling, or
only at low redshift.
\section{Modified gravity in the quasi-static limit}
\label{qsl}
Linear perturbations in MG models are often studied in the
\textit{quasi-static} (QS) limit, obtained for wavenumbers $k$ that
are large compared to the inverse sound horizon, and in which the
additional degrees of freedom with respect to GR (e.g. a scalar field
or a second spin-2 field) do not propagate significantly but rather can be well described
by constraint equations (such as the Poisson equation). The existence
of a valid QS approximation is not always guaranteed and should be
checked on a case by case basis. For instance, it may happen that
the scales at which the limit can be attained are well outside the
linear regime, or it may occur that the validity of the limit depends
on the specific initial conditions for the perturbations. In addition,
the QS limit has to be properly defined in order to avoid spurious
scale dependencies \cite{Bellini2014}.
Working with a perturbed flat FLRW metric in Newtonian gauge
\begin{equation}
\mathrm{d}s^{2}=-(1+2\Psi)\mathrm{d}t^{2}+a^{2}(1+2\Phi)\mathrm{d}x^{i}\mathrm{d}x^{i}\,,
\label{met}
\end{equation}
the effects of MG in the QS limit can be generally encoded
in two functions \cite{Amendola:2012ky} that reduce to $Y=\eta=1$
in $\Lambda$CDM):
\begin{equation}
Y(k,a)\equiv-\frac{2k^{2}\Psi}{3{\cal H}^2 \Omega_{m}\delta_{m}}\,,\quad\eta(k,a)\equiv-\frac{\Phi}{\Psi}\,,\label{Y_def}
\end{equation}
where ${\cal H}=aH$ is the conformal Hubble function, $\delta_{m}$ is the matter density contrast and $\Omega_{m}$
is the background matter energy density relative to the critical
one.
In eq.(\ref{Y_def}) the perturbation variables are meant to denote the standard deviations of the respective quantities; therefore $Y,\eta$ are deterministic functions.
The function $Y$ gives an indication of how the growth
of matter perturbations is altered with respect to the standard one
in GR for large $k$. The function $\eta$ depends
on the two Newtonian gravitational potentials and therefore effectively
on the anisotropic stress.
It has been shown that $Y$ and $\eta$ take a particularly simple form in broad classes of MG models in which the equations of
motion for the perturbations are of second order (for instance, in the Horndeski Lagrangian or in bimetric gravity)
\cite{DeFelice:2011hq,Amendola:2012ky,2014arXiv1402.1988K,2014arXiv1404.4061S}:
\begin{align}
Y=h_{1}\frac{1+(k/{\cal H})^2 h_{5}}{1+(k/{\cal H})^{2}h_{3}}\,,\quad\eta=h_{2}\frac{1+(k/{\cal H})^2 h_{4}}{1+(k/{\cal H})^{2}h_{5}}\,.\label{limo1}
\end{align}
Here $h_{1-5}$ are functions that depend only on time and can be
obtained directly from the Lagrangian of the model (see Appendix). Several types of Hordenski Lagrangians
that have been studied in detail (see e.g. references in
\cite{Bellini2014}) provide specific examples for which the expressions
(\ref{limo1}) can be applied.
As discussed in the Introduction, we will focus our attention on models
such that $Y=1$ but $\eta\neq1$. With this choice we ensure that
scalar perturbations obey the standard Poisson equation for large
$k$: therefore on scales below the sound horizon, the growth of scalar
perturbations follows the usual one in $\Lambda$CDM. This can occur if
$h_{1}=1$ and $h_{3}\simeq h_{5}$ (see Appendix \ref{appe}). To
simplify our task further, we will also assume that the three remaining
$h$-functions ($h_{2}$, $h_{4}$ and $h_{5}$) can be treated as
constants. This amounts to say that their time variation is slow in
one Hubble time and it is a reasonable approximation for studying
first order scale dependent effects. With this simplification, a very
large class of MG models is mapped into three real constants that
encode the possible effects of the anisotropic stress.
As crude as they may seem, these approximations are a significant
improvement with respect to earlier studies of linear perturbations
in MG, in which $Y$ and $\eta$ were often assumed to be pure constants.
The lensing effect of MG can then be easily described by the deviation
of $\eta$ with respect to unity, which we parametrize as follows:
\begin{align}
1+\eta=2a_{1}\frac{1+a_{2}(k/k_{p})^{2}}{1+a_{3}(k/k_{p})^{2}}\,,\label{def_sigma}
\end{align}
where $a_{1},a_{2},a_{3}$ will be assumed to be constant and we will
take $k_{p}=0.1h$/Mpc.%
\footnote{As usual, $h$ here is the reduced Hubble constant $h=H_{0}/100$,
where $H_{0}$ is the present rate of expansion expressed
in km/s/Mpc.%
}\\
\section{Tensor modes propagation speed}
\label{sec:speed} If the anisotropic stress is non-standard (i.e.
$\eta\not=1$), it can be shown that the tensor modes propagate in
a non-standard way. As discussed earlier, this will modify the CMB
spectra and, in particular, the B--modes. The general form of the
propagation equation for the transverse-traceless amplitude $h$ in Hordenski Lagrangians can be written (see \cite{DeFelice:2011bh,Bellini2014}
and also \cite{Saltas:2014dha} for a generalization)
as:\footnote{We thank M. Kunz, I. Saltas and I. Sawicki for discussing with us the structure of this equation.}
\begin{equation}
\ddot{h}+(3+\alpha_{M})H\dot{h}+c_{T}^{2}\frac{k^{2}}{a^{2}}h=0\,\,\,,\,\label{eq:gw}
\end{equation}
where dots represent derivatives with respect to cosmic time and $\alpha_{M},c_{T}$ are functions of time that depend
on the specific model; in the standard case one has $\alpha_{M}=0$
and the gravitational waves speed $c_{T}$ equals the speed of light,
$c_{T}=1$. Notice that the tensor equation is valid in general, not
just in the QS limit. In the notation of \cite{Amendola:2012ky,Bellini2014}
one has:
\begin{align}
\alpha_{M} & =\dot{w}_{1}/w_{1}H\,\,\,,\\
c_{T}^{2} & =w_{4}/w_{1}\,\,\,,
\end{align}
where $w_{1},w_{4}$ are in general time-dependent functions explicitly
defined in the Appendix that depends on the MG model. Although in
principle they can be both non-zero, in the specific case we are investigating
here (see Appendix), one has $\alpha_{M}=0$ and
\begin{equation}
c_{T}^{2}=w_{1}=\frac{1}{2a_1-1}\,\,\,.\label{gwct}
\end{equation}
A decrease of $c_{T}$ moves the horizon crossing of tensor modes
to later times and smaller scales; as a consequence, the BB spectrum
tensor mode first peak moves to higher $\ell$s, as we show later
on. The position of the tensor B peak is therefore a measure of the
gravitational speed at decoupling time.
The speed of gravitational waves can be constrained also with the
gravi-Cherenkov effect (see e.g. \cite{1980AnPhy.125...35C, Moore:2001bv,2012JCAP...07..050K}), which gives an extremely tight lower limit but no upper limit. Other possible ways to constrain the graviton speed are reviewed in \cite{Goldhaber:2008xy}. However, all these methods apply only locally (or at most within the distance scale of cosmic rays) and/or at the present time; therefore, they are complementary to the observation of B-modes. For other recent analysis on quantum gravity effects see for example \cite{2014arXiv1404.6672C}.
The theoretical BB spectrum shows another peak at $\ell\approx5$,
still to be detected, due to the effects of tensor modes on the scattering
during reionization. Also this peak gets shifted for $c_{T}\not=1$
as we show below. Its detection, for instance with the proposed satellite mission LiteBIRD \footnote{http://litebird.jp/} \cite{2013arXiv1311.2847M}, could therefore put constraints on
the gravitational wave speed before and during reionization.
\section{Weak lensing and CMB spectra}
\label{lensCMB} In order to compute the weak lensing of the CMB by
LSS in a flat universe, we define the lensing potential $\psi$ from
the Weyl potential $\tilde{\Psi}\equiv (\Psi - \Phi)/2$ as follows \cite{Lewis_Challinor_2006} (note that we are using a different signature for the metric in eq. (\ref{met})):
\begin{align}
\psi(\mathbf{\hat{n}})= 2 \int_{0}^{\chi_{\ast}}d\chi\,\tilde{\Psi}(\chi\mathbf{\hat{n}},\tau_{0}-\chi)\,\frac{\chi_{\ast}-\chi}{\chi_{\ast}\chi}\,,\label{lens}
\end{align}
In this expression, $\mathbf{\hat{n}}$ is a unit vector in three-space
that gives the (non-deflected) direction of propagation of a CMB photon,
$\tau_{0}-\chi$ is the conformal time at which the photon was at position
$\chi\mathbf{\hat{n}}$ and $\chi_{\ast}$ is the conformal distance
to the last scattering surface (assuming it is instantaneous).
The deflection angle with respect to the trajectory that the photon would have in a perfectly
homogeneous universe is given by the angular derivative of the lensing
potential $\alpha = \nabla_{\mathbf{\hat{n}}}\psi(\mathbf{\hat{n}})$. The
lensed CMB temperature $\Theta_{L}$ measured in the direction $\mathbf{\hat{n}}$
corresponds to the unlensed temperature $\Theta$ in the direction
$\mathbf{\hat{n}}+\nabla_{\mathbf{\hat{n}}}\psi$, i.e.
\begin{align}
\Theta_{L}(\mathbf{\hat{n}})=\Theta\left(\mathbf{\hat{n}}+\nabla_{\mathbf{\hat{n}}}\psi\right)\,.
\end{align}
With the definitions (\ref{Y_def}) we can write
\begin{equation}
\tilde{\Psi}=\frac{1}{2}(1+\eta)\Psi
\label{psieq}
\end{equation}
and use this relation to express the lensing potential (\ref{lens})
in terms of $\Psi$ and $\eta$. We can then write $P_{\tilde{\Psi}} = (1+\eta)^2 P_{\Psi}/4$ where the power spectrum (and similarly the transfer function) of $\tilde{\Psi}$ is written in terms of the one of $\Psi$, which is related to the
matter one by the Poisson equation (\ref{Y_def}).
As noticed before, since the main contribution to CMB lensing comes
from a short range in $z$, peaked around $z\sim1$ \cite{2004PhRvD..70b3515A,acquaviva_baccigalupi_2006,2013JCAP...09..004C, 2013JCAP...02..024A},
the assumption of constant values for $a_{1,2,3}$ is justified. We can then modify a Boltzmann code to include the effect of
MG in the computation of the lensing potential with these approximations.
The power spectrum $C_{l}^{\psi}$ of the lensing potential at a
given $\mathbf{\hat{n}}$ is
\begin{align}
\langle\psi_{lm}\psi_{l'm'}\rangle=\delta_{ll'}\delta_{mm'}C_{l}^{\psi}\,,
\end{align}
where $\psi_{lm}$ are the coefficients of the expansion of $\psi$
in spherical harmonics: $\psi=\sum_{lm}\psi_{lm}Y_{lm}$. This power
spectrum can be expressed as \cite{Lewis_Challinor_2006}:
\begin{equation}
C_{\ell}^{\psi}=16\pi\int{\frac{dk}{k}\mathcal{P}_{\mathcal{R}}(k)\left[\int_{0}^{\chi_{\ast}}d\chi\, T_{\tilde{\Psi}}(k;\tau_{0}-\chi)\, j_{l}(k\chi)\,\frac{\chi_{\ast}-\chi}{\chi_{\ast}\chi}\right]^{2}}\,,\label{cll}
\end{equation}
where
${\cal{P}}_{\tilde{\Psi}} = \mathcal{P}_{\mathcal{R}} T_{\tilde{\Psi}}$ and $j_{\ell}(r)=\sqrt{\pi/2r}J_{l+1/2}(r)$, being $J_{l}(r)$ is the $l$--th Bessel function of the first kind. The transfer function
$T_{\tilde{\Psi}}(k;\tau_{0}-\chi)$ propagates the lensing potential (\ref{lens})
forward in time. $\mathcal{P}_{\mathcal{R}}(k)$
denotes the power spectrum of the primordial curvature perturbation
$\mathcal{R}$ at the last scattering surface.
Following eq.(\ref{psieq}), we can then account for the effect of MG by introducing a $(1+\eta)/2$ factor inside the $\chi$
integral of (\ref{cll}). Again, for $\eta = 1$, the integral is the standard one.
Before doing the full calculation with a Boltzmann code, we can get
some insight of the MG effect in the region of relevance for the expected primordial B-mode peak
for large scales $(\ell\ll1000)$ \cite{Lewis_Challinor_2006}. In this limit
and at lowest order in $C_{\ell}^{\psi}$, the lensed B-mode spectrum
$\tilde{C}_{l}^{B}$ is approximately independent of $\ell$ \cite{Lewis_Challinor_2006}:
\begin{align}
\tilde{C}_{\ell}^{B}\simeq\frac{1}{4}\int d\ell'\ell^{\prime5}C_{\ell'}^{\psi}C_{\ell'}^{E}\,,\label{bmodex}
\end{align}
where $C_{\ell}^{E}$ is the unlensed E-mode spectrum. Since $C_{\ell}^{\psi}$ enters linearly
in (\ref{bmodex}), we see that in modified gravity the lensing contribution
to the B-mode spectrum gets enhanced at large scales by $(1+\eta)^{2}/4$,
with respect to $\Lambda$CDM.
Since we are deriving the lensing MG effects in the QS limit,
i.e. at sub-horizon scales, it is necessary to check the consistency
of our assumption. A comoving mode $k$ translates through the Limber
approximation into a multipole
\begin{equation}
\ell\approx\pi r(z)k\,.
\end{equation}
One finds that for
a standard $\Lambda$CDM background, $k/{\cal H}>10$ up to $z=7$, and $k/{\cal H}>5$
up to $z=20$, assuming $\ell\ge100$, which is where most of the
BB lensing signal is expected. Since the lensing effect comes from
$z$ of a few at most \cite{Lewis_Challinor_2006}\textbf{, }we expect
the quasi-static approximation to be acceptable.
As we already mentioned, we are neglecting the impact
of MG on the integrated Sachs-Wolfe (ISW) effect. Although $Y=1$
means that there is no effect on the matter perturbation growth, there
will be a change in the ISW due to the fact that photons see both
potentials $\Psi,\Phi$, just as in the lensing case. However, contrary
to lensing, the ISW has a non negligible impact on the temperature
CMB fluctuations at superhorizon scales, where our quasi-static approximation
is not reliable. We can obtain a rough estimate of the MG effect on
the low-$\ell$ TT spectrum by considering that the ISW contribution
is negligible at $\ell>30$. Then we find that it amounts to less than 20\% (see e.g. \cite{2004A&A...423..811M})
at $\ell>10$, and rises up to 50\% for $\ell=2$. The MG effects
are again proportional to $(1+\eta)^{2}$, although at these
scales the form of $\eta$ is no longer given by the QS expression.
Nevertheless, assuming $a_1\approx1.3$ (see our best fit case below),
we see that the ISW should be increased by 69\%, which means the total
TT spectrum should increase by 14\% at $\ell\approx10$, and up to
35\% at the quadrupole. This is a non negligible effect but it is
well within the cosmic variance. Moreover, it might be possible to absorb
it by slight adjustments of other cosmological parameters, although we are not
going to explore this possibility in depth in this paper.
Let us just briefly point out what would be the effect of the MG model we consider on the determination of neutrino masses and the total number of relativistic species. As it is well known (see e.g. \cite{Hou:2012xq}), the effect of small neutrino masses on the CMB power spectrum takes place via secondary anisotropies and is multipole dependent. In standard $\Lambda$CDM, a total neutrino mass of the order of $\sum m_\nu\simeq 0.5$ eV reduces the CMB $C_l$ with respect to the case $\sum m_\nu =0$ by at most $\sim 8\%$ for $10< \ell <20$, due to the late time ISW. For larger values of $\ell$, the early time ISW decreases the $C_l$ by smaller amount (around $2\%$ up to $\ell \sim 100$) and then increases it by approximately $1\%$ (for $100<\ell <500$). As we have just explained the $(1+\eta)^2$ factor introduced in the ISW by the modification of gravity we consider in this work tends to enhance the $C_l$ with respect to $\Lambda$CDM for all $\ell$. Therefore, for small multipoles ($\ell < 100$) it would go on the opposite direction as that of $\sum m_\nu$, whereas it would enhance the effect of massive neutrinos for $100 < \ell <500$ (where the early time ISW introduced by $\sum m_\nu$ is rather small). If a $(1+\eta)^2$ factor was present in the CMB, the net overall effect would be an increased ISW that could be approximately compensated by a higher value of $\sum m_\nu$.
On the other hand, the main CMB effect of a larger number of effective relativistic species $N_{\text{eff}}$ with respect to its standard value of 3.046 is due to a delayed time of matter-radiation equality, which results in a vertical shift of the CMB peaks with respect to the first one \cite{Hou:2011ec}. Concretely, it introduces a shift of the order of $\Delta C_l/C_l\simeq -0.072\, \Delta N_{\text{eff}}$. This means that a $(1+\eta)^2$ ISW change can have a similar effect as that of a reduced number of relativistic species. In conclusion, the kind of MG that we consider can affect the determination of both the total neutrino mass and the number of relativistic species. A detailed study of the degeneracies of these effects could be the object of future work.
In any case, in order not to bias our results, we decided to bypass
this problem in the data analysis below by cutting all the multipoles
$\ell<100$ in the TT and TE spectra. This slightly enlarges our errors
but avoids using the incorrectly modeled low-$\ell$ tail of the temperature
spectrum.
We conclude this paragraph by noting that also weak lensing of the matter power spectrum can be used to test modified gravity models \cite{2013MNRAS.429.2249S} and should be seen as a complementary probe with respect to the lensing on B-mode polarization. On the other hand, we have checked that the corresponding TT and EE spectra are very little affected by the kind of modification of gravity that we study,
so that
the main contribution of MG is on B-modes.
\section{Results}
\label{results}
In order to test how modified gravity affects the CMB spectra, we have
modified the publicly available Boltzmann code CAMB.%
\footnote{http://camb.info/%
} Given the assumptions described in \ref{qsl}, we do not need to
modify the background, which is assumed to be $\Lambda$CDM. Within
CAMB, we apply two sets of corrections to perturbations, which can
be activated separately or simultaneously:
\begin{itemize}
\item we modify the lensing potential as described in eq.~(\ref{psieq}), which
then depends on the $a_{1},a_{2},a_{3}$ parameters;
\item we modify the tensor propagation equation as described eq.~(\ref{eq:gw}). This modification only depends
on $a_{1}$.
\end{itemize}
We have included the three new parameters in COSMOMC \cite{cosmomc_lewis_bridle_2002}.
Both CAMB and COSMOMC used for this paper are the ones from the March
2014 version. After checking that for $a_{1}=1$ and $a_{2}=a_{3}$
we recover standard $\Lambda$CDM, we have then tested separately
three cases: when MG effects are present at all times and therefore
both on lensing and on the gravitational wave speed (we refer to this
case as ``CT+lensing''), when they are present only at decoupling
and therefore only on the tensor speed (``CT''), or only at low
redshift (``lensing'').
To illustrate the effects of MG on the BB polarization we show its
spectrum
in Figures (\ref{fig:BB_all} -
\ref{fig:BB}) for various choices of $a_{1}$ while we always fix
$a_{2}=a_{3}$, just for illustration.
\begin{figure}
\centering %
\begin{tabular}{cc}
\includegraphics[width=85mm]{BB_freec_r0_vara_lensing} & \includegraphics[width=85mm]{BB_freec_r2dn1_vara_lensing}\tabularnewline
& \tabularnewline
\includegraphics[width=85mm]{BB_freec_r2dn1_vara_CT} & \includegraphics[width=85mm]{BB_freec_r2dn1_vara_CT_lensing}\tabularnewline
\end{tabular}\caption{BB power spectrum for MG models. In the top panels we show the effect
of a lensing correction for $r=0$ and $r=0.2$ respectively. In the
bottom left panel we plot the case in which the 'CT' correction on
the speed of gravitational waves is active (while the lensing is standard).
In the bottom right panel we activate both effects. In all cases for
simplicity we assume $a_{2}=a_{3}$ and different values of the $a_{1}$
parameter. The corresponding values of $c_T^2$ are written in the bottom left panel and are related to $a_1$ via eq.(\ref{gwct}).
For comparison, the predictions for $\Lambda$CDM with $r=0$ (short
dashed, blue) and $r=0.2$ (solid, blue) are also shown. The black
dots are the data points from BICEP2. }
\label{fig:BB_all}
\end{figure}
In Fig.(\ref{fig:BB_all}) we plot the BB spectrum for the three
effects (lensing, CT, CT + lensing). The top panels refer to the lensing
case for a tensor to scalar ratio $r_{0.05}=0$ and $r_{0.05}=0.2$
respectively. The $\Lambda$CDM is also shown for reference for both
values of $r_{0.05}$. As expected, increasing the value of $a_{1}$
effectively increases $\eta$ and therefore the integrand in eq.(\ref{cll}):
the lensing contribution therefore increases in amplitude and extends
to smaller multipoles than expected for the corresponding $\Lambda$CDM.
The position of the primordial peak is however not affected (although
its amplitude also receives a contribution from MG). The bottom left
panel shows the BB spectra when only the CT effect is active: in this
case, the speed of gravitational waves changes with the inverse of
$a_{1}$ as described in eq. (\ref{gwct}): a value of $a_{1}$ smaller
(larger) than one increases (decreases) the speed of gravitational
waves with respect to $\Lambda$CDM and shifts the expected position
of the primordial peak to smaller (larger) multipoles. When both effects are active, as in the bottom right
panel, both the lensing amplitude and the shift in peak position can
occur. In all panels we also overplot for reference the recently released
data points of BICEP2.%
\footnote{http://bicepkeck.org/; B2\_3yr\_bandpowers\_20140314.txt
} We note that the corresponding TT and EE spectra are very little
affected by these changes, so that in our model the main contribution of
MG is on B-modes.
In Figure (\ref{fig:BB_largel}) we replot the lensing case, as in
Fig. (\ref{fig:BB_all}) top left panel, for $r_{0.05}=0$ and a wider
range in scales, up to $\ell=2000$, to show how the lensing predictions
for MG compare with the available data from POLARBEAR,%
\footnote{http://lambda.gsfc.nasa.gov/product/suborbit/polarbear\_prod\_table.cfm
} although we do not use them in our analysis. Notice that if we took
BICEP2 data at face value, while the ``bump'' around $\ell\sim80$
is best rendered by a non-zero tensor-to-scalar ratio ($r=0.2$) in
$\Lambda$CDM, the location of the upper points is qualitatively in
good agreement with a modification of gravity represented by $a_1 \approx1.5$.
Finally, in Figure (\ref{fig:BB}), we zoom in the low-$\ell$ region
in order to emphasize the effect of a change in $c_{T}$ on the reionization
peak. Interestingly, future measurements of the reionization peak
by experiments like those in Refs.\cite{2012SPIE.8442E..19H,2009arXiv0906.1188B,2011JCAP...07..025K,2013arXiv1306.2259P,2011arXiv1102.2181T} could be used to discriminate MG theories. For example, the satellite mission LiteBIRD \cite{2013arXiv1311.2847M} would have a sensitivity to characterise the tensor to scalar ratio r with an uncertainty of $\delta r \sim 0.001$. At these scales the lensing contribution is negligible
and the only modification comes from the correction in the
speed of gravitational waves.
\begin{figure}[!]
\centering \includegraphics[width=0.633\textwidth]{BB_freec_r0_vara_lensing_large}
\caption{The theoretical predictions of Fig. (\ref{fig:BB_all}), first panel,
are shown here for a larger $\ell$ range with a $\log$ scale. The
horizontal error bars associated to the BICEP2 data points correspond
to the interval $(\ell_{min},\ell_{max})$ from the data currently
available from the BICEP2 collaboration. For reference, although not
used in this analysis, we also over plot data from POLARBEAR (magenta,
triangular points). The third of these points is given as an upper
limit at 2 standard deviations.}
\label{fig:BB_largel} \includegraphics[width=0.63\textwidth]{BB_freec_r2dn1_vara_CT_reio}
\caption{Zoom at low multipoles of the BB spectrum of Fig.(\ref{fig:BB_all}),
bottom left panel. A CT effect, i.e. a change in the speed of gravitational
waves, is expected to modify the reionization peak in B-modes. }
\label{fig:BB}
\end{figure}
We then proceed with Monte Carlo simulations using available data to test the parameters of
our implementation of MG. We use WMAP9 data \cite{wmap9},
ACT \cite{das_etal_2013} + SPT \cite{reichardt_etal_2012} and the data from BICEP2 \cite{Ade:2014xna} on B-modes
polarization. We enforce the inflationary consistency relation $n_t\simeq -r/8$ for the
tensor spectral index $n_{t}$. We perform different runs, including
the three cases illustrated before: lensing modification, CT, CT +
lensing modifications. Results are illustrated in Table \ref{Tab:values}.
As the values for $a_{2}$ and $a_{3}$ can in principle span several
orders of magnitude, we use for convenience the logarithm of these
quantities. The parameters $a_{2}$ and $a_{3}$ play no role in CT and
are essentially unconstrained also including the lensing MG effect. Provided that the
foreground contributions are under control, B-modes polarization will
be however a very good probe to test the $a_{1}$ parameter, i.e. the
anisotropic stress and the speed of gravitational waves.
\begin{table}[htdp]
\begin{centering}
\begin{tabular}{|c|c|c|c|}
\hline
\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,
Datasets: WMAP9 + HighL + BICEP2\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, \tabularnewline
\hline
\end{tabular}\\
%
\begin{tabular}{|c|c|c|c|c|c|}
\hline
Parameter & $\Lambda$CDM & MG (lensing) & MG (CT) & MG (CT + lensing) \tabularnewline
\hline
$a_{1}$ & - & 1.30 $\pm$ 0.16 & $2.89\pm1.21$ & $1.28\pm0.15$ \tabularnewline
\hline
$\log_{10}\, a_{2}$ & - & < 0.30 & - & $<-1.12$ \tabularnewline
\hline
$\log_{10}\, a_{3}$ & - & < 0.37 & - & $<-0.11$ \tabularnewline
\hline
$r_{0.05}$ & 0.21 $\pm$ 0.05 & 0.19 $\pm$ 0.05 & $0.35\pm0.11$ & $0.21\pm0.05$ \tabularnewline
\hline
$r_{0.002}$ & 0.23 $\pm$ 0.06 & 0.21 $\pm$ 0.06 & $0.43\pm0.17$ & $0.23\pm0.07$ \tabularnewline
\hline
$H_{0}$ & 72.0 $\pm$ 2.2 & 73.7 $\pm$ 2.4 & $73.0\pm2.4$ & $73.9\pm2.4$ \tabularnewline
\hline
$n_{s}$ & 0.998 $\pm$ 0.013 & 1.000 $\pm$ 0.014 & $1.012\pm0.016$ & $1.006\pm0.015$ \tabularnewline
\hline
\hline
$-\log\mathcal{L}$ & 4146 & 4142 & 4145 & 4143 \tabularnewline
\hline
\end{tabular}
\par\end{centering}
\caption{Mean values $\pm$ standard deviation for a selection of parameters.
Both columns refer to the combination of datasets WMAP + HighL + BICEP2.
For $\log(a_{2})$ and $\log(a_{3})$ parameters we write the 95$\%$
upper limit; these parameters are essentially not constrained. }
\label{Tab:values}
\end{table}
In Fig.(\ref{fig:1d_abc}) we show 1D posterior contours for a selection
of cosmological parameters and different runs. Here it becomes clearer
that $a_{1}$ is mainly constrained by the lensing contribution, while
CT gives much larger uncertainty. This can be seen intuitively from
Fig.(\ref{fig:BB_all}): CT influences both the amplitude and the
position of the BB spectrum and is able to fit the data for a larger
range in $a_{1}$.
In order to check the validity of the QS limit, we test
the effect of the incorrectly modelled low-$\ell$ part of the TT
spectrum by redoing the MCMC simulation cutting the first 100 multipoles
in the TT and TE spectra. A comparison of the run with and without cut
is shown in Fig.(\ref{fig:1d_abc_cut}). As we see, while parameters
like $n_{s}$ and $H_{0}$ are affected by the cut, constraints on
$a_{1}$ do not depend on the cut; i.e. they
do not depend much on the low $\ell$ multipoles in the temperature
spectra. This is reassuring as it shows that the QS limit
and its simplest numerical implementation may be sufficient to test
MG theories.
\begin{figure}
\begin{centering}
\includegraphics[width=0.8\textwidth]{mg_wmap_highL_selection_abc_CT}
\caption{One-dimensional posterior contours for a selection of cosmological
parameters. We compare the three cases in which: lensing is modified
(solid, black line), CT is modified (dotted magenta line), CT and
lensing are both modified (light solid, green line). We over plot
also the case of $\Lambda$CDM for comparison (dot dashed, blue line).
}
\label{fig:1d_abc}
\par\end{centering}
\centering{}
\end{figure}
\begin{figure}
\begin{centering}
\includegraphics[width=0.8\textwidth]{mg_wmap_highL_selection_abc_cut}
\caption{One-dimensional posterior contours for a selection of cosmological
parameters. For the case in which both CT and lensing are both modified,
we compare the constraints obtained using all multipoles (solid, green
line), to the case in which we cut all multipoles $\ell<100$ in TT and TE (dash-dotted, light green).}
\label{fig:1d_abc_cut}
\par\end{centering}
\centering{}
\end{figure}
In Fig.(\ref{Fig:2d}) we show the comparison of the 2D posterior
contours for the three effects. In addition to the
considerations already done above, we notice here that $a_{2}$
and $a_{3}$ are degenerate and tend to align along the direction
$a_{2}=a_{3}$. This particular direction removes any scale dependence in
the expression for $\eta$ in (\ref{def_sigma}).
\begin{figure}[htp]
\centering %
\begin{tabular}{ccc}
\includegraphics[width=55mm]{compare_mg_r005_wmap_BICEP_highL_2D_abc_bc} &
\includegraphics[width=55mm]{compare_mg_r005_wmap_BICEP_highL_2D_abc_ar} &
\includegraphics[width=55mm]{compare_mg_r005_wmap_BICEP_highL_2D_abc_rn}\tabularnewline
\end{tabular}\caption{2D posterior contours for a selection of cosmological parameters and
the three cases in which lensing only is modified (green contours),
CT only is modified (orange/light contours), CT and lensing are both
modified (blue/darker contours). The case including CT
only does not depend on $a_{2}$ and $a_{3}$ parameters. As before,
we consider the data combination WMAP + HighL+ BICEP2.}
\label{Fig:2d}
\end{figure}
In Fig.(\ref{fig:BB_largel_bestMG}) we redo Fig.(\ref{fig:BB_largel})
for $a_1=1.30$, corresponding to the mean (and best fit) of modified gravity, for the lensing case shown in the plot.
\begin{figure}
\begin{centering}
\includegraphics[width=0.6\textwidth]{BB_bestMG_largel_color3_r005}
\caption{Best fits from Table 1. MG with $r\neq0$ is shown (dash-dotted, orange)
together with $\Lambda$CDM assuming also $r\neq0$ (dashed, blue).
For comparison we also show $\Lambda$CDM with $r=0$ (blue). As in
Figure 2, we show the data points of POLARBEAR (magenta triangles)
in addition to those of BICEP2. The error in the abscissa associated
to the BICEP2 data points corresponds to the interval $(\ell_{min},\ell_{max})$
from the release by the BICEP2 team. The third point from POLARBEAR
is plotted as an upper limit at 2 standard deviations. We recall that
POLARBEAR data are not used in the analysis and are only shown here
for reference.}
\label{fig:BB_largel_bestMG}
\par\end{centering}
\begin{centering}
\par\end{centering}
\centering{}
\end{figure}
Finally, we remap the constraints obtained for the various runs on
$a_{1}$ into $c_{T}$, the speed of gravitational waves. The resulting
1D contours are shown in Fig.(\ref{Fig:ct_remapped}). We find that
$c_{T}=0.8\pm 0.07$ from CT+lensing. Using CT alone the constraint is much weaker: $c_T \gtrsim 0.4$ at 2$\sigma$.
The reason for this behavior can be understood by looking at Fig. \ref{Fig:2d}. In the central panel one can see
that $a_1$ (or equivalently $c_T$) is quite degenerate with $r_{0.05}$ if lensing is not taken into account. This is because a shift of the tensor peak towards higher $\ell$s can be partially
compensated by an increase of $r_{0.05}$. In other words, $a_1$ (or $c_T$) could be determined by
the tensor peak position (see the bottom-left panel of Fig. 1) which is not firmly established by the current data. However,
$c_T$ changes the lensing amplitude in a significant way (see Fig. \ref{fig:BB_all}, top-right panel) and is therefore well measured by lensing alone.
\begin{figure}[htp]
\centering %
\begin{tabular}{ccc}
\includegraphics[width=95mm]{mg_wmap_highL_abc_CT_remapped} & & \tabularnewline
\end{tabular}\caption{Speed of gravitational waves, as obtained remapping the constraints
on $a_{1}$ for the three different effects considered in this paper:
lensing (dark solid, black line), CT (dotted magenta), CT + lensing
(lighter solid, green). $c_{T}^2=1$ corresponds to the standard value.}
\label{Fig:ct_remapped}
\end{figure}
\section{Conclusions}
\label{conc}
The polarized light from the last scattering surface carries important
information in addition to the temperature anisotropy. Not only it
helps constraining the cosmological parameters but it also allows
to separate the effects of primordial gravitational waves from scalar
perturbations, both predicted by inflationary models. Current experiments
\cite{Ade:2013gez,Ade:2014xna, ActPol2014} are already providing
results which will soon be cross tested by Planck. Future observations \cite{2012SPIE.8442E..19H,2009arXiv0906.1188B,2011JCAP...07..025K,2013arXiv1306.2259P,2011arXiv1102.2181T}
will keep improving our knowledge
of CMB polarization. B-modes, once foregrounds are accounted for, are particularly important in this context
since they contain both the imprint of primordial gravitational waves
and the one from gravitational lensing induced by large-scale structure. In
a sense, B-modes are the ultimate test of gravity at cosmological
scales since they probe two genuinely general relativistic
effects.
Modifications of gravity have been mainly proposed to describe the late time acceleration of the universe (see e.g. \cite{Amendola2010a} for a review) but they have also been studied as a possibility for driving primordial inflation (e.g. \cite{Kobayashi:2011nu, Gao:2011qe, DeFelice:2011uc, Tsujikawa:2013ila}). In any case, it is conceivable that some extra degrees of freedom affect gravity at early times, see for instance \cite{Konnig:2014xva}. In this paper we investigated how B-modes can be employed to test gravity at early times and at cosmological scales. Both lensing and gravitational
wave propagation depend on the features of gravity and one can use them to constrain its properties. It is remarkable that both effects
depend on the amount of anisotropic stress
$\eta$, which in general differs from the standard value of unity in
modified gravity. To simplify our study, we single out the effects
that depend on $\eta$ alone by selecting models in which the
background and the matter perturbation growth are exactly standard.
Moreover, we work in the quasi-static limit, in which all
the modified gravity effects can be embodied in just two arbitrary
functions of time and space. In a vast class of models (the Horndeski Lagrangian \cite{Horndeski:1974} and
in bimetric gravity \cite{Hassan:2011zd}), these two functions
take a particularly simple form, given by eq.(\ref{limo1}).
We show by a Monte Carlo analysis with real data that it is
indeed possible to constrain the anisotropic stress and the gravitational
wave speed with B-modes. Although the particular values we obtain
here are to be taken with great caution because the B-mode available data are
still under close scrutiny, future data has great potential for providing tight constraints on MG.\footnote{Shortly after our work was made public, a related study on the speed of gravitational waves also appeared, see \cite{Raveri:2014eea}.}
Several
of the assumptions we adopted for simplicity in this work can be lifted relatively
easily: one can for instance remove the assumption of constant
MG parameters and work with the full equations rather than with their quasi
static limit. This will be addressed in future works.
\begin{acknowledgments}
We acknowledge support from the TRR33 - The Dark Universe - DFG Grant.
We thank Emilio Bellini, Diego Blas, Alicia Bueno-Belloso, Martin Kunz, Ippocratis Saltas,
Ignacy Sawicki, and Licia Verde for fruitful discussions. We also thank Francesco Montanari for useful comments on a draft version of this work.
\end{acknowledgments}
|
\section{Discussion \& Conclusion}
The issue of stability is relevant for similarities induced by topic models using approximate inference techniques. The correlations between similarities from identical but randomly initialized models, can be used as a tool to gain some insight into this matter. From the preliminary results on the music example we find the induced similarities (fig. \ref{fig:stability}) to be highly stable.
Furthermore, inspecting the similarities obtained from different data types; figure \ref{fig:cross}, we observe that while the audio model in itself does not seem to provide higher intra- than inter-genre similarity, it is still significantly positively correlated to the other modality group which does possess some discriminative power in terms of genre labels. Moreover, it seems that an increasing number of topics causes the correlation between similarities from models estimated on different modality groups to decrease. We speculate that this is linked to the specific topic model variant, for which \cite{Blei2003annotated} also note that the model describes the joint distribution of different modalities well, but does not model the relations between them.
In conclusion, we have proposed the multi-modal LDA as a method to define similarities in \hyphenation{multi-media}multimedia applications with multiple heterogeneous data sources based on the predictive-likelihood. This was extended with the Mantel test allowing direct evaluation of the consistency and correspondence of the resulting similarities.
\section{Introduction}
Calculating similarity between objects linked to multiple data sources is more urgent than ever. A prime example is the typical multimedia application of music services where users face a virtually infinite pool of songs to choose from. Here choices are based on many different information sources including the audio/sound, meta-data like genre, and social influences \cite{salganik2006experimental}, hence, attempts of modeling the geometry of music navigation have taken on a multi-modal perspective.
In fusing heterogeneous modalities like audio, genre, and user generated tags it is both a challenge to establish a combined model in a 'symmetric' manner so that one modality do not dominate others and it is challenging to evaluate the quality of the resulting geometric representation. Here, we focus on the latter issue by testing the consistency of derived inter-song (dis-)similarity by means of direct comparison between similarities using the Mantel permutation test.
Topic models have previously been used to infer geometry in the image and music domain, e.g. by \cite{Yoshii2006hybridplsa} combining audio features and listening histories. In \cite{Lienhart2009a} images and tags were analyzed, also by means of a multi-modal topic model. In \cite{Hoffman2008musichdp} music similarity is inferred with a nonparametric Bayesian model, and \cite{Blei2003annotated} describe multiple multi-modal extensions to basic LDA models and evaluate the models on an image information retrieval task. Furthermore, topic model induced similarities among documents have been put to use in a navigation application \cite{Chaney2003visualizingtm}, and different similarity estimates are also discussed in relation to a content-based image retrieval problem \cite{Horster2007}.
\section{Model \& Inference}
\vspace{-0.1cm}
\begin{figure}[t]
\centering
\begin{subfigure}[b]{0.62\textwidth}
\begin{mdframed}
\begin{small}
\begin{itemize}[leftmargin=*]
\item For each topic indexed by $t\in[1;T]$ in each modality indexed by $m\in[1;M]$\\
Draw $\bm\phi_t^{(m)}\sim \textit{Dirichlet}(\bm\beta^{(m)})$\\
This is the parameters of the $t^{th}$ topic's distribution over vocabulary $[1;V^{(m)}]$ of modality $m$.
\item For each document indexed by $d\in[1;D]$
\begin{itemize}
\item Draw $\bm\theta_d \sim \textit{Dirichlet}(\bm\alpha)$ \\
This is the parameters of the $d^{th}$ documents's distribution over topics $[1;T]$.
\item For each modality $m\in[1;M]$
\begin{itemize}
\item For each word $w$ in the $m^{th}$ modality of document $d$
\begin{itemize}
\item Draw a specific topic $z^{(m)}\sim \textit{Categorical}(\bm\theta_d)$
\item Draw a word $w^{(m)} \sim \textit{Categorical}(\bm\phi_{z^{(m)}}^{(m)})$
\end{itemize}
\end{itemize}
\end{itemize}
\end{itemize}
\end{small}
\end{mdframed}
\caption{Generative process}
\label{fig:minipage1}
\end{subfigure}
\quad
\begin{subfigure}[b]{0.32\textwidth}
\centering
\includegraphics[width=1\textwidth, angle=90]{mmLDAmodel.eps}
\caption{The multi-modal Latent Dirichlet Allocation model represented as a probabilistic graphical model.}
\end{subfigure}
\caption{}
\label{fig:mmLDA_model}
\vspace{-1.3em}
\end{figure}
To be able to measure similarities between objects, a representation of these objects is needed. In this work we use a version of Latent Dirichlet Allocation that incorporates multiple sources of information into a joint object representation similar to \cite{Blei2003annotated}. In \cite{polyling2009Mimno}, this model was applied to a multilingual corpus.\\
Each object is represented by a multinomial distribution over topics which is common for all of the modalities composing the object. Each topic is defined by a set of multinomial distributions over features, each of which is defined on the vocabulary specific for a modality. To explain the characteristics of the model, the assumed generative process for objects is outlined in figure \ref{fig:mmLDA_model} together with a graphical representation of the model. The difference from a number of individual LDA models, each defined on a separate modality, is that each object is described by a single, shared distribution over topics, which potentially induces strong dependencies between the feature distributions representing the same topic in the individual modalities.
Performing inference in the model amounts to estimation of the posterior distributions over the latent variables. We use a Gibbs sampler inspired by the sparsity improvements proposed by \cite{sparseLDA-Yao}. For evaluation (see section \ref{sec:result}), we use point estimates $\bm\theta^s$ and $\bm\phi^s$ derived from a sample $\mathbf z^s$ from the Markov chain, by taking the expectations of the respective posterior Dirichlet distributions defined by $\mathbf z^s$. In this work we choose the state of the chain with the highest model evidence within the last 50 out of 4000 iterations. Hyper-parameters are optimized using fixed point updates \cite{minka_polya,wallach_phd_thesis}. The prior on the document topic distributions is an asymmetric Dirichlet with parameter $\bm\alpha$, and the priors over the vocabularies of the respective modalities are symmetric Dirichlet distributions with parameters $\bm\beta^{(m)}$
\subsubsection*{Acknowledgment}
\vspace{-3.5mm}
This work was supported in part by the Danish Council for Strategic Research of the Danish Agency for Science Technology and Innovation under the CoSound project, case number 11-115328. This publication only reflects the authors' views.
\bibliographystyle{unsrt}
\renewcommand\refname{\vskip -0.7cm}
\subsubsection*{References}
\small
\section{Experimental Results: Music Similarity}\label{sec:result}
In this preliminary study we examine induced similarities in a subset of the Million Song Dataset~\cite{Bertin-Mahieux2011}, consisting of 30.000 tracks with equal proportions of 15 different genres. Each track is composed of data from a number of different sources:
Open vocabulary tags from users (last.fm),
Lyrics (musiXmatch.com),
Editorial artist tags (allmusic.com),
Artist tags (musicBrainz),
User listening history (echonest),
Genre and style (allmusic), and
Audio Features (echonest).
All modalities---besides the audio features---are naturally occurring as counts of words and for the audio we turn to an \textit{audio word} approach, where the continuous features are vector quantized into a total of 2144 words.
For this pilot study we estimate topic models on combinations of groups of modalities from the mentioned list, respectively consisting of the first 5, the genre and style labels, and the audio. To be able to assess the model stability of the similarities, we estimate each model five times from different random initialisations of the Markov chain. This is done for every training set of a 10-fold cross-validation split. The correlations between all combinations of the 5 similarity matrices resulting from each held-out fold are then calculated, and the resulting distributions of correlation coefficients are shown in figure \ref{fig:self_corr}. Figure \ref{fig:meta_cross} shows the distributions of correlations between similarities based on audio and on the larger modality group. The correlations are evidently much smaller than for identical models, but a Mantel test with 100 permutations suggest that the null hypothesis of no correlation can be rejected at a significance level of at least 1\% for all three model complexities.
\begin{figure}[b]
\centering
\begin{subfigure}[b]{0.41\textwidth}
\includegraphics[width=1\textwidth]{selfCorr_spear.eps}
\caption{{\small Boxplots of Spearman's correlation between similarity matrices obtained using the same parameters and data, but different random initialisation. Each row correspond to a modality group.}}
\label{fig:self_corr}
\end{subfigure}
\qquad
\begin{subfigure}[b]{0.41\textwidth}
\includegraphics[width=1\textwidth]{meta_scatter.eps}
\caption{{\small Scatter plot of two specific similarities obtained by random initialisations of the same model estimated from the larger group of modalities. The colors indicate within- and between-genre similarities.}}
\end{subfigure}
\caption{}
\label{fig:stability}
\end{figure}
\begin{figure}[t]
\centering
\begin{subfigure}[b]{0.41\textwidth}
\includegraphics[width=1\textwidth]{artisttermVSaudio_corr.eps}
\caption{{\small Boxplots of Spearman's correlation between similarity matrices obtained using the larger modality group and the audio.}}
\label{fig:meta_cross}
\end{subfigure}
\qquad
\begin{subfigure}[b]{0.41\textwidth}
\includegraphics[width=1\textwidth]{audio_meta_scatter.eps}
\caption{{\small Scatter plot of the two examples of similarities obtained from topic models with same parameters, but two mutually exclusive modalities of data.}}
\label{fig:meta_cross_scatter}
\end{subfigure}
\caption{}
\label{fig:cross}
\end{figure}
\section{Similarities in Topic Models}
\vspace{-0.1cm}
As already hinted, there are many ways to define and calculate similarities in topic models; both between topics and documents. In this paper we focus on the latter. Most methods in literature are based solely on the distributions of topics in the documents, $\bm\theta$, e.g. \cite{Hoffman2008musichdp} measures the Kullback-Leibler divergence between two such distributions, while \cite{Horster2007} also mentions inner products and cosine similarities as candidates. With focus on visualization, \cite{Chaney2003visualizingtm}, introduces the yet another dissimilarity measure based on topic proportions. \cite{Horster2007} promotes a measure based on the predictive likelihood of the document contents, and this approach is the basis of the method chosen here;
The similarity of two documents $A$ and $B$ is given by the mean per-word log-likelihood of the words of document $A$ given the topic distribution of document $B$ (and the vocabulary distributions).
\begin{align}
\frac{\log p(\mathbf w_A|\bm\theta_B^s,\bm\phi^s)}{\sum_{m=1}^MN^{(m)}_A}, \text{\,\,\,\,where\,\,\,\,} p(\mathbf w_A|\bm\theta_B^s,\bm\phi^s)=\prod_{m=1}^M\prod_{i=1}^{N_A^{(m)}}\sum_{t=1}^T{(\bm\phi^{(m)}_{t,w^{(m)}_{Ai}})}^{\top}\bm\theta_{t,B}
\end{align}
We use this approach to calculate a non-symmetric similarity matrix between all objects in the held-out cross-validation fold, for which the topic proportions have been estimated using ``fold-in''.
\footnote{For the few held-out documents that do not contain any words in the modalities used for model estimation, we chose to simulate a uniform distribution of words in such an empty document by one occurrence of every word in the vocabulary.}
While this similarity measure is more computationally demanding than e.g. the KL-divergence, when the number of topics $T$ used in the model increases, it might happen that some topics have vocabulary distributions that are very alike and only differ on a few words. Thus two documents with mainly the same type of content may have large proportions of different topics, causing them to be very dissimilar according to a topic proportion based measure. For a non-parametric topic model such as \cite{Hoffman2008musichdp}, this might not be a large concern, however, for parametric topic models, this should be taken into consideration. Generally, most of the discussed similarity measures are not proper metrics in the geometric sense, but for (dis-)similarity purposes the exact properties might not be important, depending on the application.
\textbf{Comparing Similarities - the Mantel test}\\
An important aspect of this work is the ability to assess the relations between different similarities induced by models estimated from multiple, possibly different, heterogeneous data sources. To compare such similarities we look at the correlation between the defined similarities. For testing the significance of the correlations we can apply a Mantel style test \cite{Mantel1967test}. The Mantel test is a non-parametric test to assess the relation between two (dis-)similarity matrices.
The null hypothesis is that the two matrices are unrelated, and the null distribution is approximated by calculating the test statistic for a large number of random permutations of the two matrices (excluding the diagonal elements); permuting rows and columns together to maintain the distribution of (dis-)similarities for each object. In this work we use Spearman's correlation coefficient as the test statistic.
|
\section{Definition and validation}
The ExF is a node property derived from local network topology, independent of the rest of the network or any specific spreading process. It is formally defined as follows. Consider a network with one infected node $i$ and all remaining nodes susceptible.
Enumerate all possible clusters $J=1, \ldots, |J|$ of infected nodes after $x$ transmission events, assuming no recovery (See Fig 1).
Generally speaking, $x=2$ is sufficient and assumed for the rest of this manuscript.
Hence $J$ includes all possible combinations of $i$ plus two nodes at distance one from $i$, and $i$ plus one node at distance one and one at distance two.
The enumeration is over all possible orderings of the transmission events.
Two neighbors of the seed ($a$ and $b$) form two clusters
($[i \rightarrow a, i \rightarrow b$] and $[i \rightarrow b, i \rightarrow a$])
or, if $a$ and $b$ also share an edge, four clusters.
Define the degree of a cluster of nodes $d_j$ as the number of edges connecting nodes inside the cluster to nodes outside.
After two transmissions without recovery, the FoI of a spreading process seeded from node $i$ is a discrete random variable taking a value in $(d_1, \ldots, d_J)$, allowing for the proportionality constant equal to the transmission rate of the process.
The \emph{expected} FoI can be approximated by the entropy of the $d_j$ after normalisation
\begin{equation}
ExF(i)=- \sum_{j=1}^{J} \bar{d_j} \log(\bar{d_j})
\end{equation}
where $i$ refers to the seed node and $\bar{d_k}=\frac{d_k}{\Sigma_Jd_j},\, \forall k \in J$.
The entropy is needed for generating the expected value due to the extreme variability in the shape, number of modes, and number of terms in the distributions of $d_j$ for different seed nodes.
Setting $x=2$ is recommended but not required.
Increasing the number of transmissions beyond two adds very little information while increasing the computational cost
\stepcounter{note}(Note S\arabic{note}), in agreement with other proposed spreading power metrics~\cite{Bauer2012,Klemm2012} and
consistent with the decaying influence of longer paths in the calculations of the eigenvalue, subgraph, and related centralities \cite{Bonacich1987,Estrada2005,Estrada2010, Benzi2013}.
In certain cases, however, it may be desirable to consider more transmission events. For example, a node at the end of a chain of length two can only form one transmission cluster of size two, hence its ExF is zero.
Comparing two such nodes requires setting $x=3$, in which case a subscript can be used for clarity (e.g. ExF$_3$).
One modification may be in order for SIS/SIR processes, inspired by the following. Imagine a node with degree one connected to a hub. While such a node will have a high ExF, its chance of realizing this force depends entirely on transmitting to the hub before recovery.
Such nodes are common in dense social networks. For example, 84\% of the 225K nodes in an EU institution email network~\cite{Leskovec2007} have degree one.
In such networks,
it may be helpful to
multiply the ExF by the log of the seed node's degree
after first rescaling the seed's degree by some factor $\alpha>1$.
\begin{equation}
ExF^{M}(i)= \log(\alpha\, deg(i))\, ExF(i)
\end{equation}
The rescaling is motivated in that the log of one is zero, and the ExF$^M$ is most informative in networks where many nodes have degree one.
The rescaling factor must be greater than one, and should also be small to avoid overpowering the influence of the degree.
In the rest of this manuscript, we use $\alpha=2$, the smallest integer which satisfies these criteria.
Computing the ExF$^M$ for $\alpha$ ranging from 1.0001 to 16 does not substantively alter the metric, as all such variations show correlations greater than 0.99 to ExF$^M$ computed with $\alpha=2$ \stepcounter{note}(Note S\arabic{note}).
Running times are discussed in \stepcounter{note} Note S\arabic{note}.
Example code providing an implementation of the ExF is available at\\
https://github.com/glennlawyer/ExpectedForce.
\begin{figure}[t]
\centerline{\includegraphics[width=.45\textwidth]{onwardcon.png}}
\caption{This network will be in one of eight possible states after two transmissions from the the seed node (red). Two states are illustrated, where the seed has transmitted to the two orange nodes along the solid black edges. Each state has an associated number of (dashed orange) edges to susceptible nodes (blue), the cluster degree. States containing two neighbors of the seed can form in two ways or, if they are part of a triangle, four ways. The eight network states associated with the pictured seed node arrise from thirteen possible transmission clusters. The ExF of the seed node is the entropy of the distribution of the normalized cluster degree over all possible transmission clusters.}
\end{figure}
\subsection{Correlation to epidemic outcomes}
We measure correlations between ExF and epidemic outcomes
on five families of simulated networks chosen such that their densities and degree distributions span a wide range of human contact structures (Methods, and Table S4).
One hundred random networks of 1,000 nodes are generated in each family. Further comparison is made using a suite of twenty four empirical networks ranging from 1,133 to 855,800 nodes (Table S5).
Epidemic outcome
are observed by simulating multiple epidemics in both continuous and discrete time from a number of seed nodes in each network.
Correlations are measured between these outcomes and the ExF, ExF$^M$,
accessibility, eigenvalue centrality, and the k-shell of the seed nodes
(Note S3).
\begin{figure*}[t]
\begin{center}
\centerline{\includegraphics[width=.8\textwidth]{Figure-2-lawyer.png}}
\caption{Observed correlations between metrics and epidemic outcomes in each network family. The ExF and ExF$^M$ (orange shades) are consistently strong, with mean correlations greater than 0.85 and small variance. The other measures (k-shell, eigenvalue centrality, and accessibility, blue-green shades) show both lower mean values and higher variance. Each violin summarizes correlations computed on 100 simulated networks. Spreading processes (x axis) are suffixed to indicate simulations in continuous (-C) or discrete (-D) time. }
\end{center}
\end{figure*}
The ExF is highly predictive of all epidemic outcomes on all networks tested, simulated and empirical. Mean correlation to SI process outcomes is 83\% on simulated and 74\% on empirical networks. For processes with recovery, mean correlation is 91\% on simulated and 82\% on empirical networks. Standard deviations over the one hundred simulated networks in each family are typically $0.02-0.03$. The 95\% confidence bounds on empirical networks are in the same range.
In almost all cases the ExF significantly outperforms the accessibility, eigenvalue centrality, and the k-shell (difference in mean correlations greater than the standard deviation of the higher mean).
The ExF is only bested in the following three cases, each time by the k-shell: discrete time SIR processes in the simulated Astrophysics (0.97 vs 0.92) and Facebook (0.97 vs 0.94) networks and the SI process on the ``email-EUAll'' network (0.50 vs 0.41).
The performance of the k-shell was surprisingly strong, given that two previous studies by independent groups have observed rather poor performance for this metric~\cite{Klemm2012,daSilva2012}.
The observed correlations on 100 simulated networks in each family are shown in Fig. 2 (also Table S6).
The measured correlations and their standard errors for all empirical networks are shown in Fig. 3 (also Tables S7-9).
The ExF's predictive power is robust to variation in network structure.
The theory behind the ExF$^M$ suggests that the ExF might perform less well for SIS/SIR processes on denser networks, yet
mean correlation for continuous time SIS processes is barely changed between the loose Pareto/Amazon networks (0.93/0.95) and the dense Astrophysics/Facebook networks (0.92/0.90).
As expected, the predictive power of the ExF$^M$ improves on the denser networks
(mean correlations: Pareto/Amazon 0.89/0.92, Astrophysics/Facebook 0.94/0.95).
The accuracy of the accessibility metric, in contrast, collapses for all spreading processes on the dense networks (mean correlation over all spreading processes: Pareto/Amazon 0.74/0.90, Astrophysics/Facebook 0.28/0.20.)
A previous analysis which observed similar poor performance for the accessibility on dense networks concluded that spreading processes seeded from nodes with low accessibility are not capable of entering the epidemic phase~\cite{daSilva2012}.
Our results show this is not the case, as these nodes have a small yet observable epidemic potential which the ExF is able to capture and quantify.
Performance of the k-shell and the eigenvalue centrality is also strongly influenced by network structure. For SIS/SIR processes, both showed higher mean and sharply reduced variance on the denser networks.
In an interesting contrast, the k-shell's predictive power for SI processes is reduced in denser networks.
The eigenvalue centrality's performance also varies by spreading process, showing its best performance on discrete time SIS models-- though again this variation is modulated by network density.
Two other independent groups have observed that relationships between centrality rankings and epidemic outcomes are strongly influenced by network structure and the parameters of the spreading processes~\cite{Sikic2013,Bauer2012},
leading the authors of \cite{Sikic2013} to conclude that these measures severely underestimate the epidemic impact of structurally peripheral nodes.
\subsection{Weighted graphs}
\hspace{1ex} The ExF generalizes to graphs with weighted edges, where we assume the edge weights correspond to per-edge transmission likelihoods. Use these weights to calculate the probability of each way that each cluster could occur, and re-define the cluster degree as the sum of all edge weights leading out from that cluster. The extension to directed graphs is also straightforward; limit the
enumeration to edges leading from an infected to a susceptible node.
We test this generalization by computing the weighted and unweighted ExF on 1,000 node networks with Pareto (1,2.3) degree distributions and edge weights chosen according to one of the following
three distributions:
uniformly distributed between one and three,
uniformly distributed between one and ten,
and exponentially distributed with unit rate, weights rounded up to the nearest integer.
Fifty networks were simulated for each distribution of edge weights.
Correlation between the weighted and unweighted ExF was greater than 0.99 for all network edge weighting distributions tested.
As expected from the tight correlation, the weighted and unweighted ExF showed no meaningful difference in predictive ability, which remained high.
Observed correlations between node ExF and epidemic potential in discrete-time SIS processes were 0.88/0.89 $\pm 0.03$ (unweighted/weighted ExF) under the uniform-3 scheme,
0.83/0.04 $\pm 0.03$ under the uniform-10 scheme, and
0.80/0.79 $\pm 0.05$ under the exponentially distributed weighting scheme.
\begin{figure*}[t]
\begin{center}
\centerline{\includegraphics[width=.8\textwidth]{Figure-3-lawyer.png}}
\caption{Observed correlations and 95\% confidence intervals between metrics and epidemic outcomes on the 24 empirical networks. The ExF and ExF$^M$ (orange shades) show strong performance, consistently outperforming the other metrics (k-shell, eigenvalue centrality, and accessibility when computed, blue-green shades). The suffix ``-D'' indicates spreading processes simulated in discrete time. }
\end{center}
\end{figure*}
\section{Discussion}
The ExF predicts all types of epidemic outcomes with high accuracy over a broad range of network structures and spreading processes. The low variance in observed correlations over multiple simulated network and epidemic models shows that the measure is robust, as do the tight confidence bounds on empirical networks.
What, then, does it tell us about the nature of node spreading power?
The definition of the ExF implies that spreading power is determined by both the degree of the node and the degree of its neighbors, and that the relative influence of these two factors is different for nodes of low versus high spreading power.
Weaker nodes gain what strength they have from their neighbors onward connectivity, whereas more influential nodes get their strength from their large number of connections.
These relationships are accentuated by network density.
This is a result of the combinatorics behind the enumeration over transmission clusters.
Clusters are formed either by choosing two neighbors from the first geodisc, a term which grows approximately quadratricly in node degree, or by each path leading to the second geodisc, a term which grows approximately linearly in the size of this disk.
Generally speaking, the second geodisc is much larger than the first.
Hence spreading power is determined largely by terms from the second geodisc until the quadratic growth of the influence of the first overtakes it.
The influence of network density reflects the ExF's sensitivity to
network motifs such as triangles and squares.
Each triangle touching the seed node contributes four terms to the enumeration $J$,
meaning the contribution of the first geodisc grows faster than quadratically.
Sensitivity to squares and triangles is also granted by the entropy term.
These motifs, which are more common towards the cores of dense networks, reduce the disparity of the cluster degree distributions thus increasing the ExF
(Fig. 4).
\begin{figure}[t]
\begin{center}
\centerline{\includegraphics[width=.4\textwidth]{Figure-4-lawyer.png}}
\caption{Plotting ExF (x-axis) versus node degree (orange), the sum of the degree of all neighbors (blue), and the sum of the degree of all neighbors at distance 2 (green) shows that for nodes with low ExF, the neighbor's degree has strong correlation to ExF, while for nodes with high ExF their own degree is more closely correlated ways. The result is accentuated in denser collaboration networks in comparison to more diffuse Pareto networks.}
\end{center}
\end{figure}
The approach taken by the ExF is fundamentally different than that taken by most centrality measures.
Centrality measures typically set out to produce a ranking which identifies the most influential nodes in the network, under the assumption that highly influential nodes are those with the maximal sum of some type of walk~\cite{Bauer2012,Borgatti2005,Borgatti2006,Estrada2010,Benzi2013}.
The choice of the appropriate type, scaling, and length of walks
contain implicit assumptions regarding network flows \cite{Borgatti2005}, cohesion structure~\cite{Borgatti2006}, and/or other topological characteristics of the network~\cite{Estrada2010,Benzi2013}.
The k-shell is a slight exception, as it was originally intended to precipitate out the most cohesive regions of the network rather than to explicitly rank nodes within cohesive regions~\cite{Seidman1983}, yet it is now recognized as one of the best centrality measures for identifying a network's most influential spreaders~\cite{kitsak2010}.
Spreading power metrics generalize the walk counting framework by explicitly including transmission probabilities when scaling the walks \cite{Bauer2012,Klemm2012,Travencolo2008,Viana2012}.
The question not asked is if the type, scaling, and lengths of walks best suited to identifying the most important nodes applies equally well to the rest of the network.
To the extent that the optimal choice of factors depends on network topology, then the difference in topology between core and periphery suggests that choices well suited to the core are seldom appropriate for the remainder of the network.
Both the combinatorics behind the ExF and the walk counting behind most centrality measures agree that
influential nodes are those which combine high degree with a preponderance of influential neighbors.
The ExF has high rank correlation with both the eigenvalue centrality and the k-shell (0.62-0.92 across the simulated network families \stepcounter{note} (Note S\arabic{note} and Table S2).
Likewise, the ExF has 60-90\% agreement with the eigenvalue centrality on the top ten network nodes and 100\% agreement with the k-shell.
The difference between walk counting and the ExF is that the ExF adapts the relative influence of different walks and walk lengths based on local connectivity, whereas approaches based on functions of the adjacency matrix apply a fixed protocol.
The eigenvalue centrality is weighted node degree, where the weights are the importance of the neighbors \cite{Bonacich1987,Estrada2005}.
But the eigenvalue centrality is strictly a global measure, unable to distinguish more subtle variations in local structure \cite{Estrada2005,Benzi2013}.
The k-shell erodes node degree to match the number of neighbors with similar degree. Since this discards remaining information on the individual degree of nodes within a common shell, the accuracy of its predictions is heavily influenced by the number of shells in the network.
The accessibility combines node and neighbor degree into a measure of the number of nodes likely to be reached by walks of a given length~\cite{Viana2012}.
But this approach has difficulties quantifying nodes in dense, small diameter networks, which accentuate differences between core and peripheral topology.
Calculation of the ExF depends only on the local topology.
This allows epidemic outcomes on the whole network to be predicted with high accuracy even when only a small portion of the network is known \stepcounter{note}(Note S\arabic{note}).
It is rare for the full structure of a real network to be fully known; typically the network structure is inferred from indirect, incomplete, and often biased observations.
Specification of an adjacency matrix is even more difficult when the underlying network is dynamic.
In contrast,
estimates of eigenvalue centrality fluctuate depending on which nodes are sampled~\cite{Costenbader2003}. Both the pagerank~\cite{Ghoshal2011} and the k-shell~\cite{Adiga2013} are highly sensitive to pertubations in network topology, making them unreliable for incomplete or noisy systems.
Reliance on a local network does, however, lead to one weakness in the ExF.
A network may contain large but disparate communities.
Here, a node serving as a bridge between two communities
might be able to spread a process to the entire network with more force than a node far from the bridge, even when the second node has more (local) spreading power than the bridging node.
The ExF's local nature makes it blind to these larger topological constraints on spread.
The ExF is strongly correlated to epidemic outcome, outperforming existing metrics of node spreading power and centrality.
The measure depends only on local network topology, allowing its use in dynamic as well as static networks.
For most nodes, the most important determinant of their spreading power is the sum of their neighbors' degree. As node power grows, so does the importance of the node's own degree. This relationship is accentuated in denser networks.
\section{Methods}
\subsection{Epidemic simulations and outcomes}
The epidemic outcome for SI process is the time to half coverage (\textit{tthc}); the time until exactly half the nodes in the network are infected.
This is measured at each seed node \stepcounter{note}(Note S\arabic{note}) by simulating 100 spreading processes and fitting the observed \emph{tthc} from all simulations to a gamma distribution.
Simulations are run in continuous time, with the time of the next transmission event drawn from an exponential distribution with rate equal to the current number of infected-susceptible edges.
Discrete time does not provide sufficient resolution for SI processes.
Spreading processes with recovery are of interest when the ratio of transmissibility to recovery for the process $\beta$ is in the critical range which allows for but does not guarantee an epidemic.
The epidemic outcome is a node's epidemic potential (\textit{EPo}), the probability that a given node can seed an epidemic \stepcounter{note}(Note S\arabic{note}).
The \textit{EPo} is measured by simulating 100 outbreaks per seed node and counting the number which result in at least half of the network nodes becoming infected.
Continous-time simulations model the time of the next transmission as in the SI model just described,
except the transmission rate parameter scales the current number of infected-susceptile edges by some $\beta$ in the critical range (see below),
and the time of the next recovery from a unit rate exponential distribution weighted by the number of infected individuals.
In each round of the discrete-time simulations, transmission occurs along each infected-susceptible edge with probability $r=-\log(1-\beta)$ (standard conversion of rate to probability) and nodes recovering at the end of the round.
The critical range for $\beta$ can be defined empirically using the criterion that if $\beta$ is in the critical range, then a large majority of nodes will have \textit{EPo} $\in [2\%, 98\%]$.
We here set $\beta$ independently for each network to a fixed multiple of $1/\lambda$, where $\lambda$ is the largest eigenvalue of the adjacency matrix.
Similarity in network structure allows the same multiple to be used for both the Pareto and Amazon networks, likewise the Astrophysics and Facebook networks.
Internet hyperlinks lie between these two other classes, and the multiples from the social networks yields good results.
This proceedure resulted in at least 96\% of nodes in continuous-time simulations with \emph{EPo} in the critical range; in discrete time the relevant figure is 76\% or better (Table S3).
\subsection{The networks}
The five families of simulated networks are defined by their degree distributions,
one theoretical (Pareto), and four derived from the following empirical human contact networks:
the Amazon co-purchase network from May 2003 ~\cite{Leskovec2007a},
the graph of Internet hyperlinks from the 2002 Google programming contest~\cite{Leskovec2009},
the collaboration network from ArXiv Astrophysics between 1993 and 2003~\cite{Leskovec2007}, and
Facebook wall posts from the New Orleans network~\cite{viswanath2009} (Table S4).
The Pareto and Amazon networks are characterized by large diameter and low density.
The Astrophysics and Facebook networks are two orders of magnitude more dense and have correspondingly smaller diameter.
Google's map of Internet hyperlinks lies in between the other two families.
The networks can also be characterized by the largest eigenvalue of their adjacency matrix, as theory suggests that the critical disease transmission probability dividing epidemic from extinction regimes is the inverse of this value~\cite{Klemm2012}, which further implies that a network's inherent susceptibility to disease spreading is reflected by this eigenvalue.
Again, the selected network families cover a wide range of inherent susceptibility to disease.
Simulations are conducted using giant components of 1,000 nodes.
Networks with a Pareto (1,2.3) degree distribution are simulated using the Chung Lu protocol, a versatile and general method for constructing random graphs with given expected degree sequence~\cite{Chung2002}.
The remaining networks are simulated by randomly sampling 1,000 values from the degree sequence of the actual graph without replacement and generating a graph from these values using the Igraph function ``degree.sequence.game'' which generates undirected, connected simple graphs matching the input degree sequence \cite{igraph}.
Empirical networks were selected and downloaded from the Stanford Large Network Repository (SNAP) and Alex Arenas's collection according to the following criteria:
having between 1,000 and 1,000,000 nodes in the largest connencted component,
representing one clear network structure,
and, in the case that the same network is sampled at multiple timepoints, the latest timepoint.
Twenty one networks from SNAP and two from Alex Arena passed these criteria.
The simulated Amazon networks are derived from the earliest Amazon co-purchase network in SNAP.
For completeness, this network is also included in the suite of empirical networks.
The Facebook network was downloaded from the Max Planck Institute for Software Systems.
For the purpose of testing, networks are treated as undirected graphs with multiple and self-edges removed.
The 24 networks are characterized in Table S5, which includes the internet address of the collections.
The size of the empirical networks required slight modifications to the overall approach.
Seed nodes are 1,000 nodes selected uniformly at random.
Epidemics with recovery are simulated only in discrete time.
As can be seen from the results on the random networks, discrete time simulations provide approximately the same mean outcome as continuous time, and only slightly higher variance.
The transmission/recovery probability ratio $\beta$ is determined independently for each network by binary search over possible values until one is found such that a minimum of 80\% of tested nodes have \emph{EPo} between 0.05 and 0.95.
When the network has more than 25,000 nodes, the \textit{tthc} is measured as the time of the 1,000$^{th}$ transmission rather than the time when half the network is infected.
Finally, the R software package does not gracefully handle the multiplication of matrices larger than 25Kx25K, even with the Matrix package~\cite{Bates2012}.
Hence the accessibility was not computed for networks with more than 25k nodes.
\begin{acknowledgments}
This work was supported out of the general budget of the Max Planck Society.
\end{acknowledgments}
\bibliographystyle{pnas}
|
\section{Vacuum stability analysis}
With the discovery at the LHC of a new resonance \cite{higgsdiscovery} with
mass around 125-126 GeV
and properties very compatible to those of the Standard Model (SM) Higgs boson
the complete particle spectrum of the SM is now known. The first run of the
LHC has delivered two important messages: i) no signal of
physics beyond the SM (BSM) was discovered. ii) The Higgs boson
was found where predicted by the SM.
\begin{figure}
$$\includegraphics[height=0.5\textwidth]{plotCLTEV}$$
\caption{Probability density function for the Higgs mass obtained combining
the indirect information coming from precision physics with the direct search
results from LEP and Tevatron. Courtesy of S. Di Vita.}
\label{fig1}
\end{figure}
In fig.\ref{fig1} I show the probability density function for the SM Higgs
boson mass obtained combining the information from precision measurements with
the results of the Higgs search experiments, the latter expressed in terms of
the likelihood of the search experiment normalized to the no-signal
case \cite{DaDe}. In the figure only the experimental results from LEP
and Tevatron before the turning on of the LHC are used. As shown from the
figure the SM had a sharp prediction: the mass of the Higgs boson had to be
between 114 and 160 GeV. Indeed the Higgs boson was found by ATLAS and CMS
exactly in that interval.
The fact that all the parameters of the SM have been
now experimentally determined constrains tightly the model and possibly BSM
physics. New Physics (NP), if exists, should be of the decoupling type, i.e.
it should have a marginal effect on the SM electroweak fit without spoiling its
very good agreement with the experimental results. This fact, together with
the negative result of the run I of the LHC, may put some doubts on the
expectation that NP has to be ``around the corner'', i.e. within the reach of
the LHC. In this situation it is natural to ask where the scale of NP is, or
if it can be as large as the Planck scale, $\mpl$, implying that
the validity of the SM can be extended up to $\mpl$.
One approach to answer this question is to study the
stability of the SM vacuum, or if the electroweak (EW) minimum we
live in is the true minimum of the SM effective potential, i.e. the
radiatively corrected scalar potential. The effective potential, in first
approximation, has the the same form as the tree-level one but with running
parameters ($\mu$ is the renormalization scale)
\begin{equation}
V^{\rm eff} \approx -\frac12 m^2 (\mu) \phi^2 (\mu)
+ \lambda(\mu) \phi^{4} (\mu) \sim \lambda(\mu) \phi^{4} (\mu)~,
\label{eq:1}
\end{equation}
then if we are looking at large values of the Higgs field, $\phi(\mu)
\gg v$ where $v$ is the EW minimum, the dominant contribution to the
potential is from the quartic term.
The search for the scale where $V^{\rm eff}$ becomes smaller than its
value at the EW minimum, i.e. the instability scale $\Lambda_I$, can
be replaced, given the steepness of the potential around that point, by
looking for the scale where $V^{\rm eff} =0$ or, for large values of the field,
where
$\lambda(\mu)=0$. The Higgs quartic coupling is special among the SM
couplings. Indeed $\lambda$ is the only SM coupling that is allowed to
change sign during the Renormalization Group (RG) evolution because it
is not multiplicatively renormalized. For all other SM coupling the
$\beta$ functions are proportional to their respective couplings and
crossing zero is not possible. In fact, one finds for $ \beta_\lambda \equiv
d \lambda/d \ln \mu$ at the one loop level
\begin{equation}
\beta_\lambda = \frac1{16 \pi^2} \left[
+ 24 \lambda^2
+ \lambda \left( 4 N_c Y_t - 9 g^2 -3 g^{\prime 2} \right)
{\color{red} -} 2 N_c Y_t ^4+ \frac98 g^4 + \frac38 g^{\prime 4} + \frac34
g^2 g^{\prime 2} \right]
\label{eq:2}
\end{equation}
where $N_c =3$ is the color factor of $SU(3)_c$, $Y_t$ the top Yukawa coupling
and $g$ and $g^\prime$ the $SU(2)_L$ and $U(1)_Y$ gauge couplings, respectively.
In the r.h.s. of eq. \ref{eq:2} the part not proportional to $\lambda$ contains
the top Yukawa coupling at the fourth power and with a negative sign. Thus,
for small values of $\lambda$ this is the term dominating $\beta_\lambda$ and
$\lambda$ is going to evolve towards smaller values eventually crossing
zero.
\begin{figure}
\includegraphics[width=0.49\textwidth, height=0.49\textwidth]{SMRGE1}
\includegraphics[width=0.49\textwidth]{lambda}
\begin{center}
\caption{\label{fig:run} Left: Evolution of the SM gauge couplings
$g_1=\sqrt{5/3}g_Y, g_2, ~g_3$, of the top, bottom couplings ($y_t$, $y_b$)
of the Higgs quartic coupling $\lambda$ and of the Higgs mass parameter $m$.
All parameters are defined in the $\overline{\mbox{\sc ms}}$ scheme.
Right: Zoom on the evolution of the Higgs quartic,
with uncertainties in $M_t$, $\alpha_s$ and $M_h$ as
indicated. Plots taken from ref.\cite{us}}
\label{fig:2}
\end{center}
\end{figure}
In fig.\ref{fig:2} (left) the evolution in the SM of the gauge, Yukawa and
scalar
couplings is shown. The running of the various couplings has been determined
using the state-of-the-art computations, i.e. three-loop beta
functions \cite{beta3} and two-loop matching conditions \cite{match2,us}.
The three gauge couplings and the top Yukawa
coupling remain perturbative and are fairly weak at high energy,
becoming roughly equal, within 10\%, around a scale of about
$10^{16}$~GeV. It is amusing to note
that the ordering of the coupling constants at low energy is
completely overturned at high energy with the (GUT normalized)
hypercharge coupling $g_1= \sqrt{5/3} g_Y$ being the largest coupling.
The evolution of $\lambda$ is zoomed in the
right part of fig.\ref{fig:2}. The Higgs quartic coupling remains weak in the
entire energy domain below $\mpl $. It decreases with energy crossing
$\lambda =0$, for the central values of top mass, $M_t$, the strong coupling,
$\alpha_s$, and the Higgs mass, $M_h$, at a scale of about $10^{10}$~GeV.
The fact that $\lambda$ becomes negative at a scale lower than $\mpl$
is a signal that the effective potential is unstable, i.e. at high scale is
either not bounded from below or it develops a second minimum that can be
deeper than the EW one. In both cases the idea that the SM can be
considered a valid theory up to $\mpl$ is in trouble because $v$ is no longer
the true minimum of the potential and there is a tunnelling probability
between the false vacuum $v$ and the true vacuum at high field values. However,
we can infer that NP must appear below $\Lambda_I$ to cure
the instability of the SM potential only if the lifetime of EW
vacuum is shorter than the life of the universe.
The rate of quantum tunnelling out of
the EW vacuum, given by the probability $d\wp/dV\, dt$ of nucleating a
bubble of true vacuum within a space volume $dV$ and time interval
$dt$, was first computed in the late seventies by S.~Coleman \cite{cole}.
The total probability $\wp$ for vacuum decay to have occurred during the
history of the universe can be computed by integrating $d\wp/dV\, dt$ over
the space-time volume of our past light-cone, or
\begin{equation}
\wp \sim \tau_U^4 \Lambda_B^4\, e^{-S(\Lambda_B)}~~~~~~~~~~~~~~~~~~~
S(\Lambda_B)=\frac{8\pi^2}{3|\lambda(\Lambda_B)|}.
\label{eq:3}
\end{equation}
where $\tau_U$ is the age
of the universe and $S(\Lambda_B)$ is the action of the bounce of size
$R=\Lambda_B^{-1}$.
$\Lambda_B$ is determined as the scale at
which $\Lambda_B^4 e^{-S(\Lambda_B)}$ is maximized \cite{IRS}.
In practice this
roughly amounts to minimizing $\lambda(\Lambda_B)$, which corresponds
to the condition $\beta_\lambda (\Lambda_B)=0$. By numerical inspection of
$\wp$ in eq.(\ref{eq:3}) one finds that the exponential suppression wins over
the large 4-volume factor if $|\lambda(\Lambda_B)|$ is less than $\sim 0.05$.
Fig.\ref{fig:2} shows that $\lambda$ in its RG evolution towards $\mpl$
does become negative but never too negative. In fact the running of $\lambda$
is slowing down at high energy because its $\beta$ function at high scale
becomes very small, vanishing close to $\mpl$.
At the Planck scale one then finds \cite{us}
\begin{eqnarray}
\lambda( \mpl ) &=& -0.0113 +
0.0029\left( \frac{M_h }{\rm GeV}-125.66 \right)
-0.0065 \left( \frac{M_t }{\rm GeV} -173.10 \right)
\nonumber \\
&& +0.0018 \left( \frac{ \alpha_s(M_Z) -0.1184}{\rm 0.0007} \right)
\end{eqnarray}
that implies that our vacuum is metastable, i.e. $\wp$
is extremely small (less than $10^{-100}$) or the lifetime of the EW vacuum is
extremely long much larger than $\tau_U$.
\begin{figure}[t]
$$\includegraphics[width=0.45\textwidth]{SMht}\qquad
\includegraphics[width=0.46\textwidth]{EWphase}$$
\caption{\em {\bf Left}: SM phase diagram in terms of Higgs and top pole masses.
The plane is divided into regions of absolute stability, meta-stability,
instability of the SM vacuum, and non-perturbativity of the Higgs quartic
coupling. The dotted contour-lines show the instability scale
$\Lambda_I$ in GeV assuming $\alpha_s(M_Z)=0.1184$.
{\bf Right}: Zoom in the region of the preferred experimental range of $M_h$
and $M_t$ (the grey areas denote the allowed region at 1, 2, and 3$\sigma$).
Plots taken from ref.\cite{us}.
\label{fig:4}}
\end{figure}
The study of the two-loop effective potential \cite{V2} allows us to identify
the phases of the SM. They are shown in fig.\ref{fig:4} as a function of
the Higgs and top masses. The regions of stability, metastability, and
instability of the EW vacuum are shown both for a broad range of $M_h$
and $M_t$, and after zooming into the region corresponding to the measured
values. The uncertainty from $\alpha_s$ and from theoretical errors
is indicated by the dashed lines and the color shading along the
borders. Also shown are contour lines of the instability scale.
The measured values of
$M_h$ and $M_t$ appear to be rather special, in the sense that they
place the SM vacuum at the border
between stability and metastability. In the neighborhood of the
measured values of $M_h$ and $M_t$, the stability condition is well
approximated by
\begin{equation} M_h > 129.1\, {\rm GeV} + 2.0 (M_t- 173.10\, {\rm GeV}) -0.5
{\rm GeV} \frac{ \alpha_s(M_Z) -0.1184}{\rm 0.0007} \pm 0.3\, {\rm GeV} \ .
\end{equation}
Since the experimental error on the Higgs mass is already fairly small
and will be further reduced by future LHC analyses, it is becoming
more appropriate to express the stability condition in terms of the pole
top mass or
\begin{equation}
M_t < (171.53\pm 0.15\pm 0.23_{\alpha_s} \pm 0.15_{M_h})\, {\rm GeV}~.
\label{mtsta}
\end{equation}
\section{The role of the top quark}
As is clear from the right plot in fig.\ref{fig:4}, it is the exact value of
the top mass, rather than a further refined computation, the factor that can
discriminate between a stable and a metastable
EW vacuum. Fig.\ref{fig:4}, as well as the bound (\ref{mtsta}), are obtained
using as renormalized mass for the top quark the so-called pole mass
and identifying it with the average of the Tevatron, CMS and ATLAS
measurements, $M_t= 173.10 \pm 0.6$ GeV.
This identification can be disputed in two aspects. i)
From a theoretical point the concept of pole mass for a quark is not
well defined as quarks are not free asymptotic states. Furthermore
the quark pole mass is plagued with an intrinsic non-perturbative ambiguity
of the order of $\Lambda_{QCD}$ due to the so-called infrared (IR) renormalon
effects. ii) The top mass
parameter extracted by the experiments, which we call $M_t^{MC}$, is an object
that is obtained via the comparison between the kinematical reconstruction
of the top quark decay products and the Monte Carlo simulations of
the corresponding event. The latter requires a careful modeling of the
jets, missing energy, initial state radiation contributions as well as of the
hadronization part. $M_t^{MC}$ is a parameter sensitive to the on-shell region
of the top quark but it cannot be directly identified with the pole mass.
We can write generically $ M_t^{pole} = M_t^{MC} + \Delta$ with the understanding
that the error quoted by the experimental collaborations refers to
$ M_t^{MC}$ and not to $M_t^{pole}$. The point now
is what is the size of $\Delta$.
An analysis of the phase-space regions
in the top production cross-section at hadron colliders shows that the
region possibly sensitive to IR effects contributes very little to
the total rate. Then, even assuming an uncertainty of 100 \% in the
modelling of that region, the extraction of $ M_t^{MC}$ from the total rate
will be affected only at the level of $\sim 30$ MeV. Thus we can conclude
that $ M_t^{MC}$ can be interpreted as $M_t^{pole}$ within the intrinsic
ambiguity in the definition of $M_t^{pole}$, that implies
$\Delta \sim {\cal O}(\Lambda_{QCD}) \sim 250-500$ MeV \cite{mlm}.
It is well known that short distance masses, such the one defined in
the $\overline{\mbox{\sc ms}}$ scheme, do not suffer from the IR renormalon problem.
The $\overline{\mbox{\sc ms}}$ top mass, $M_t^{\overline{\mbox{\sc ms}}}$, can be extracted directly from the total
production cross section for top quark pairs $\sigma(t \bar{t} + X)$.
A recent analysis reports $M_t^{\overline{\mbox{\sc ms}}}(M_t) = 163.3 \pm 2.7$ \cite{adm}, a value
that translated
in terms of pole mass gives for $M_t^{pole}$ a central value very close to that
obtained via the decay products but a much larger error
\begin{equation}
M_t^{\overline{\mbox{\sc ms}}}(M_t) = 163.3 \pm 2.7 ~~{\rm GeV} \rightarrow M_t^{pole} = 173.3 \pm
2.8 ~~{\rm GeV.}
\end{equation}
The use of $\overline{\mbox{\sc ms}}$ masses in the EW theory requires some specification.
Differently from QCD, in the EW theory masses, as well as $v$, are
not parameters of the EW Lagrangian.
The parameters are the gauge, Yukawa and the scalar couplings. This implies that
the definition of an $\overline{\mbox{\sc ms}}$ mass is not unique: it depends upon the definition
of the vacuum. Indeed we can define the vacuum either as the minimum of
the tree-level scalar potential or as the minimum of the radiatively corrected
potential. In the first case we get an $\overline{\mbox{\sc ms}}$ mass that is gauge invariant,
but there will be large EW corrections in the relation between the pole and
the $\overline{\mbox{\sc ms}}$ mass \cite{jeg}. In the case of the top quark they are proportional
to $M_t^4$ implying that if we want to extract directly $M_t^{\overline{\mbox{\sc ms}}}(M_t)$ from
$\sigma(t \bar{t} + X)$ we have to consider the EW radiative corrected
cross section. In the second case, when the vacuum is defined via the
corrected potential, these large corrections are absent however the resulting
$\overline{\mbox{\sc ms}}$ mass is not a gauge-invariant object. Although this fact can seem quite
awkward we should remember that an $\overline{\mbox{\sc ms}}$ mass is not a physical object and
therefore gauge invariance is not a mandatory requirement.
\begin{figure}
$$\includegraphics[width=0.49\textwidth]{mt-indirect}$$
\caption{Indirect determination of the top pole mass from EW precision
observables. Courtesy of S.~Mishima and L.~Silvestrini.}
\label{fig:5}
\end{figure}
It is clear that, because $M_t^{\overline{\mbox{\sc ms}}}$ is determined with an error much larger
than that of $M_t^{pole}$ its use in the analysis of vacuum stability will
weaken the conclusion that the EW vacuum is metastable while admitting,
within $1~\sigma$ in the top mass error, the possibility of full stability.
However, we can take a different point of view: the top pole mass is the
same object that enters the EW fit and it can be predicted now that
we know the Higgs mass quite accurately. Is the $M_t^{pole}$ value obtained
from the fit compatible with the bound (\ref{mtsta})? The answer is in
fig.\ref{fig:5} where the probability density function for $M_t^{pole}$ is
shown with the dark (light) region corresponding to $1 (2)~\sigma$
interval. From the figure it is clear that values of $M_t$ around
$171$ GeV are in the tail of the distribution with a probability of
few per cent.
\section{Conclusions}
The SM is in a very good status. The value of the Higgs mass found by
ATLAS and CMS is very intriguing. It causes the SM potential to be at the
border of the stability region. The exact value of the top mass plays the
central role between the full stability or the metastability (preferred)
options. The possibility of $\lambda > 0$ up to $\mpl$ requires a
top mass value around $171$ GeV, a number not preferred by the EW fit.
Finally, the fact that our EW vacuum is metastable with a lifetime much longer
than the age of the universe does not allow us to conclude that NP must appear
at a scale lower than the Planck scale.
\acknowledgments
I would like to thank the organizers of LC13 for their kind invitation.
I am in debt with all the people I have collaborated with on this subject:
D.~Buttazzo, S.~Di Vita, J.~Elias-Mir\'o, J.R.~Espinosa, P.P. Giardino,
G.~Giudice, G.~Isidori, F.~Sala, A.~Salvio and A.~Strumia.
Interesting discussions with F.~Maltoni
and P.~Slavich are also acknowledged.
This work was partially supported by the Research Executive
Agency (REA) of the European Union under the Grant Agreement number PITN-GA-
2010-264564 (LHCPhenoNet).
|
\section{Introduction}
\label{sec:intro}
\IEEEPARstart{A}{femtocell} base station
abbreviated as femto BS or fBS
is a small BS with low transmission power and low cost.
Femtocells can be installed by the end users to enhance the cellular networking performance at home,
of which traffic is transported via an Internet backhaul such as Digital Subscriber Line~(DSL) or cable modem.
Two-tier cellular networks, consisting of a conventional macrocell network
and underlaying short-range femtocells,
have received considerable attention from industry and academia as an efficient solution
to deal with the exploding demand for wireless data communication.
The femtocell technology has an advantage over other competing indoor wireless communication
technologies, thanks to its high capacity and backward compatibility with existing cellular technologies.
The history, current status of market and technology,
research issues, and future expectation of femtocell technology
are well summarized in~\cite{jsac12andrews}.
There are mainly three strategies for a femtocell access, namely,
closed access, open access, and hybrid access strategies.
A femtocell in a closed access mode can only be accessed by
authorized femtocell users.
On the other hand, if the owner of a femtocell installs it with the
open access mode, any macrocell user might access the
femtocell. Some previous researches~\cite{iccs09lopez,globecom10xia,icc11jo}
have shown that the deployment of open access femtocells can improve the system-wide performance by transferring some of the traffic loads in congested macrocells to the femtocells.
Hybrid access mode is a compromise between closed and open access, where
an fBS allows arbitrary nearby users to
access it like open access mode but the subscribed femtocell owners can be prioritized over
unsubscribed users. The prioritization can be implemented by using various
vendor-specific mechanisms.
In 3GPP Release~8 specification~\cite{3gppTS36.3.8},
only closed and open access modes are supported for femtocells
while the hybrid access mode has been added in 3GPP Release~9 specification~\cite{3gppTS36.3.9}.
Therefore, considering both the open and hybrid access strategies is important.
In this paper, we numerically analyze and optimize the performance of both open and hybrid access femtocell
networks. We propose the load balancing schemes which properly balance the traffic loads in macrocells and femtocells. Our load balancing schemes aim at maximizing the system-wide performance,
i.e., the average throughput of the users communicating with the macrocells,
in two-tier cellular networks with open or hybrid access femtocells,
while guaranteeing some benefits of the femtocell owners
such that femtocell users can always achieve larger throughput than macrocell users.
Such an approach not only improves the macrocell user's performance via traffic offloading from macrocell to femtocell, but also promotes the deployment of femtocells via the guaranteed benefit to the femtocell owners.
In our proposed framework, we strike a balance between a macrocell and femtocells by controlling the service area of femtocells because the amount of traffic loads is dominated by the number of associated users.
In order to maximize the offloading efficiency, orthogonal deployment --- in which
the whole wireless resources are divided into two parts, one dedicated to femtocells and the other reserved for a macrocell --- is considered, and we jointly optimize the amount of resources dedicated to the femtocells and the service area of femtocells.
The optimization problem is first studied for the open access case, and then extended to the hybrid access femtocell networks, where a variable portion of intra-femtocell resources is exclusively used by femtocell owners while remaining resources are shared with macrocell users associated with the femtocell.
The contributions of this paper are summarized as follows:
\begin{enumerate}
\item We numerically analyze the performance of macrocell and femtocell users in two-tier cellular networks
where open and hybrid access fBSs are deployed in an un-planned manner.
\item Multiple essential parameters in the open and hybrid access femtocell networks are jointly optimized
to enhance the performance of both the macrocell and femtocell users.
\item Our results show that the orthogonal spectrum dedication for open and hybrid access femtocell can be more beneficial than co-channel deployment in some aspects. The conditions that femtocell deployment on dedicated channel can be preferred are also discussed.
\item It is proved that the joint optimization of the amount of dedicated resources and the service area of
femtocells is a convex optimization problem in typical environments.
\end{enumerate}
The rest of the paper is organized as follows. Section~\ref{sec:related} presents the related work. In Section~\ref{sec:system_model}, we introduce the system model and our load balancing
problem formulation in the open and hybrid access femtocell based two-tier cellular networks.
We analyze the average throughput of each type of users in Section~\ref{sec:analysis} to complete the problem formulation.
In Section~\ref{sec:optimization_open}, we obtain the optimal
parameters of open access femtocell networks and
some theorems for optimal parameter selection are discussed. By extending the
optimization framework of open access femtocell networks,
system parameters of hybrid access femtocell are also optimized
in Section~\ref{sec:optimization_hybrid}.
In Section~\ref{sec:evaluation}, our proposed schemes are evaluated based on both numerical analysis and computer simulations.
Finally, we conclude the paper in Section~\ref{sec:conclusion}.
\section{Related Work}
\label{sec:related}
Many resource allocation schemes have been proposed for the two-tier cellular networks, but most of the proposed schemes are heuristic or locally optimized schemes~\cite{coml08guvenc,globecom08choi,icc09claussen,mobihoc09sundaresan,
globecom09bai,twc10jo,icc11kaimaletu,icc11park,mobicom11arslan,icc11hatoum}.
On the other hand, some previous work conducts optimization based on
full channel information~\cite{wiopt09han,pimrc09kim,twc11cheng} or
game theoretic model~\cite{pimrc10bennis,icc11lin,icc11bennis,tvt11ko}.
Therefore, the previous researches are different from our work which optimizes
the system-wide performance based on the long-term system information
of the two-tier cellular networks where fBSs are deployed in an unplanned manner.
Our long-term parameter optimization framework is not incompatible with but complementary to
the short-term resource allocation schemes
in the sense that the long-term optimization framework can provide good guidelines
for the parameter configuration considering the system-wide average performance.
Recent papers~\cite{twc10jo,tcom09chandrasekhar} are the most relevant previous work in the literature.
Coverage control schemes have been proposed in~\cite{twc09Chandrasekhar_coverage,twc10jo}.
The authors in~\cite{twc10jo} proposes an adaptive transmit power control scheme to control
the shape and the size of femtocell coverage.
In closed access femtocell networks, which is the system model of the above mentioned papers,
the objective of coverage control is minimizing the interference leakage from fBSs to the outdoor macrocell region
while the expected service area for the subscribed users is guaranteed.
Service area adaptation in this paper is different
because the strong signal received from a femtocell is
considered a good serving signal in open and
hybrid access femtocell networks, which allow the macrocell users to access them.
In~\cite{tcom09chandrasekhar},
the authors propose a bandwidth division scheme in the two-tier cellular networks composed of the closed access femtocells. The objective and constraints of~\cite{tcom09chandrasekhar} are different from our work
since ours
utilizes the traffic offloading gain in the open and hybrid access femtocell networks.
Furthermore, we optimize some other control parameters, such as target service area of a femtocell and intra-femtocell resource dedication ratio for a femtocell owner, together with the bandwidth division ratio to enhance the system performance.
\section{System Model and Our Framework}
\label{sec:system_model}
\subsection{System Model}
\label{subsec:femto_model}
We assume a single circular macrocell region with
the radius of $D_m$ and the area of $A_m = \pi {D_m}^2$,
where a macrocell BS (mBS) is located at the center of the circular region.
Multiple
fBSs are randomly distributed within the macrocell region according to
a homogeneous Spatial Poisson Point Process~(SPPP)~\cite{book93kingman} with intensity $\lambda_f$.
In an SPPP, the number of points in a given region follows Poisson random variable
with the mean of $\lambda A$, where $\lambda$ and $A$ are the intensity of the points
and the area of the given region, respectively.
We assume that each fBS is owned by a femtocell mobile station~(fMS),
and the fBS is located at the center of the indoor circular home region
with the radius of $D_{h}$.
The fMS
of each fBS is randomly located within the
circular home region, and the indoor home area and outdoor area are partitioned with a wall.
It is possible that the fBSs following SPPP are located closer than $2D_h$ to each other thus resulting in overlapping home areas.
Although this should not happen in practice, we argue that the proposed SPPP model still well captures the reality because it is highly unlikely to have such overlapping fBSs with the practical range of $\lambda_f$ and $D_h$.
Suppose we denote by $P_{overlap}$ the probability that two or more fBSs overlap with each other. Since $P_{overlap}$ is identical to the probability that two or more fBSs exist in the area of $\pi (2 D_h)^2$, we have
\begin{eqnarray*}
P_{overlap} &=& 1 - e^{-\lambda_f \cdot 4 \pi {D_h}^2} - \lambda_f \cdot 4 \pi {D_h}^2 \cdot e^{-\lambda_f \cdot 4 \pi {D_h}^2}\\
&=& \begin{cases}
0.079, & \textrm{in US},\\
0.081, & \textrm{in Korea},
\end{cases}
\end{eqnarray*}
\noindent where the values of $\lambda_f$ and $D_h$ are obtained from {\cite{census2010_kr,census2010_us}}.
Hence, we henceforth assume that there exists only one fBS per home, i.e., the distance between any two fBSs is larger than $2D_h$.
In Section~\ref{sec:evaluation}, it will be verified that such an approach approximates typical environments with a reasonably small analytical error (less than $1$~\%) as shown in Fig.~\ref{fig:validation}.
Note that in our simulation, we generate fBSs according to SPPP and drop a new fBS located closer than $2D_h$ to any of the previously-generated fBSs.
Similarly to fBSs, macrocell users are randomly distributed according to an SPPP
with intensity $\lambda_u$ in the whole macrocell area.
Because some macrocell users can associate with a nearby femtocell
in the open and hybrid access femtocell networks,
the macrocell users are categorized into
two types,
i.e., macrocell mobile station~(mMS) and open access mobile station~(oMS).
We refer to the macrocell user who associates with mBS as mMS, while
the macrocell user who associates with an fBS thanks to the open or hybrid access
policy is referred to as oMS.
The downlink channel gain between a BS and an MS is characterized by a pathloss and fading.
When the distance between a transmitter and a receiver is $d$, the channel
gain of the link is modeled by $\Psi\left( {Zd} \right)^{ - \alpha }$, where $\alpha$ is the pathloss exponent and $Z$ represents a fixed loss
which is dependent on the type of the link.
Different $Z$ and $\alpha$ values, i.e., $Z_1$ to $Z_5$ and $\alpha_1$ to $\alpha_5$, are
defined for different types of links as shown in Table~\ref{table:pathloss}.
$\Psi \sim \exp \left( 1 \right)$ is a
Rayleigh fast fading component which has a unit average power.
We consider multiple discrete transmission rates, where the rate is adaptively determined according
to the Signal to Interference plus Noise Ratio~(SINR) value at the receiver.
Rate index $l \in [1,L]$ corresponds to the case when the SINR lies in $[\Gamma_l, \Gamma_{l+1})$, where $\Gamma_{L + 1} = \infty$.
and the spectral efficiency of
rate index $l$
is modeled as the following based on the variable rate
M-QAM transmissions:
\begin{equation}
\label{eq:rate_eq}
b_l = \log _2 \left( {1 + \frac{{\Gamma _l }}{G}} \right),
\end{equation}
where $G$ denotes Shannon Gap introduced in~\cite{tcom97goldsmith}.
The specific rate set used in the simulations are summarized in Table~\ref{table:rate_sets}.
\begin{table}
\centering
\caption{Transmission rate set}\label{table:rate_sets}
\begin{tabular}{|c||c||c|}
\hline
Rate index, $l$ & Spectral efficiency, $b_l$ & SINR region (dB) \\
& (bps/Hz) & \\
\hline\hline
1 & 0.4922 & $[-4, 0)$\\ \hline
2 & 1.3889 & $[ 0, 4)$\\ \hline
3 & 2.8962 & $[ 4, 8)$\\ \hline
4 & 4.7364 & $[ 8, 12)$\\ \hline
5 & 6.6885 & $[ 12, 16)$\\ \hline
6 & 8.6711 & $[ 16, \infty)$\\ \hline
\end{tabular}
\end{table}
Furthermore, Table~\ref{table:parameters} provides the definitions and default values
of all notations frequently used in this paper.
\begin{table*}[ht] \centering
\caption{Definition of parameters and default values}\label{table:parameters}
\begin{tabular}{|c||c||c|}
\hline
Symbol & Description & Default value \\
\hline\hline
$D_m$, $A_m$ & Radius and area of a macrocell region & $800$~m, $\pi D_m^2$\\ \hline
$D_h$, $A_h$ & Radius and area of a home region & $20$~m, $\pi D_h^2$\\ \hline
$f_c$ & Carrier frequency & $2000$~MHz\\ \hline
$W$ & System bandwidth & $5$~MHz\\ \hline
$P_N$ & Noise power density & $-174$ dBm/Hz\\ \hline
$P_m$ & Macrocell transmit power density & $46/W$ dBm/Hz\\ \hline
$P_f$ & Femtocell transmit power density & $23/W$ dBm/Hz\\ \hline
$WL$ & Wall penetration loss & $10$ dB\\ \hline
$\Psi$ & Rayleigh fading component & N/A \\ \hline
$\alpha$, $Z^{\alpha}$& Pathloss exponent and fixed loss value & See Table~\ref{table:pathloss} \\ \hline
$b_l, \Gamma_l$ & Spectral efficiency and SINR threshold for rate index $l$& See Table~\ref{table:rate_sets}\\ \hline
$\overline N_f,\lambda_f$ & Average number and intensity of fBSs in a macrocell area & $30$, $30/A_m$ \\ \hline
$\overline N_{u}, \lambda_{u}$ & Average number and intensity of macrocell users (mMSs + oMSs) & $200$, $200/A_m$\\ \hline
$\rho$ & Ratio of femtocell resources to the whole bandwidth, $\rho \in \left[0,1\right]$& N/A \\ \hline
$d_f$, $D_{\max}$ & Service radius of a femtocell region and maximum value of $d_f$, respectively & N/A \\ \hline
$x$ & Average service area of a femtocell region, $x \in \left[X_{\min},X_{\max}\right]$ & N/A \\ \hline
$\theta$ & Resource usage probability in OA/HA-Thin, $\theta \in \left[0,1\right]$& N/A \\ \hline
$\beta$ & Ratio of resources reserved for an fMS to the whole resources of a hybrid access fBS, $\beta \in \left[0,1\right]$ & N/A \\ \hline
$\overline T_f$, $\overline T_m$, $\overline T_{o}$ & Average throughput of fMS, mMS, and oMS & N/A \\ \hline
$M$ ($K$) & Required ratio between average throughput of fMS (oMS) and mMS & $10$, $1$\\ \hline
$O_{\max}$ & Maximum average outage rate allowed for an oMS & $0.15$ \\ \hline
\end{tabular}
\end{table*}
In this paper, we
consider a fully loaded network environment where the BSs always have packets to transmit.
Furthermore, we assume that the scheduler in the mBS or the open access fBS
allocates the resource blocks in a round-robin manner
so that the whole wireless resources in the cell are equally distributed to
the associated users
in a long-term.
The basic level of fairness, i.e., intra-cell resource fairness, is guaranteed
from this assumption.
In the hybrid access fBS, the scheduler reserves some amount of resources
for the fMS while the remaining resources are allocated using a round-robin manner.
We assume that the transmission power spectral densities, i.e., power per Hz,
of the mBS and fBSs are fixed as $P_m$ and $P_f$, respectively, and the noise power density is given by $P_N$.
\subsection{Our Framework}
\label{sec:modeling}
\label{subsec:updated_formulation}
\begin{figure}
\begin{center}
\includegraphics[width=0.30\textwidth]{./hybrid_resource_new.eps}
\caption{Resource dedication in hybrid access femtocells ($\rho$ and $\beta$).}
\label{fig:hybrid_model}
\end{center}
\end{figure}
In this section, we introduce our load balancing framework
for open and hybrid access femtocell networks.
As shown in Fig.~\ref{fig:hybrid_model},
our schemes divide the whole available resources into two orthogonal sub-parts, and the two parts are dedicated to the macrocell and femtocells, respectively.
We refer to the ratio of resources dedicated to the femtocells
among the whole available resources as $\rho \in \left[ {0,1} \right]$,
and we optimize $\rho$ to properly balance the traffic loads in the macro and femtocells.
We assume that the wireless resources can be divided either in time domain or frequency domain or both.
Our framework allocates the separate resources to the femtocells because
the resource separation not only limits the side effect to the existing macrocell users due to the femtocell deployment but also maximizes the offloading capability in the two-tier cellular networks by increasing the maximum cell coverage.
Though it has been generally said that open access femtocell networks prefer the co-channel deployment
option, the results of this paper show that
algorithms based on separate bandwidth can be more beneficial
in some aspects thanks to the enhanced
offloading gain and the increased flexibility to control the performance of fMSs, mMSs, and oMSs.
Appendix~\ref{sec:orthogonal} summarizes the benefits and preferred conditions of orthogonal deployment.
A hybrid access fBS allows the macrocell users to
access like open access fBS, but the intra-cell resource scheduler gives
priority to the fMS. It is different from the intra-cell resource allocation
policy of the open access femtocell where the resources are equally allocated to all the users
without distinguishing the fMS from the other macrocell users.
As shown in Fig.~\ref{fig:hybrid_model}, we assume that $\beta$ fraction of intra-femtocell resources are
dedicated to the fMS, and the remaining resources are equally shared by the
fMS and oMSs. Open access femtocell is a special case of hybrid access femtocell where $\beta=0$.
\begin{figure}
\begin{center}
\includegraphics[width=0.40\textwidth]{./ho_model_2.eps}
\caption{Femtocell service radius ($d_f$) and user associations.}
\label{fig:ho_model}
\end{center}
\end{figure}
Our load balancing scheme in the open and hybrid access femtocell networks
jointly optimizes the average service area of a femtocell as well as the
amount of bandwidth dedicated to femtocells and the amount of intra-femtocell
resource dedicated to an fMS.
We
optimize the service coverage of a femtocell because
the cell selection based on the strongest RSS value is not efficient
to promote the load balancing.
Fig.~\ref{fig:ho_model} describes the system model for the service coverage optimization.
We refer to the service radius of a femtocell as $d_f$, and the femtocell and macrocell
users who are located closer than $d_f$ associate with the femtocell rather than
the macrocell. Each fBS is required to fully cover the indoor home area, i.e., $d_f \ge D_h$.
The maximum service radius, i.e., $D_{max}$, is constrained by the physical limitation to support wireless communications. In our work, $D_{max}$ is defined by the maximum distance where the average outage probability
of an oMS is less than or equal to $O_{\max}$ while the lowest transmission rate is used.
The service radius is chosen in the range of
$d_f \in \left[ {D_h ,D_{\max } } \right]$ to properly balance the traffic loads in macro and femtocells.
The service coverage adaptation is implemented by using a simple MS initiated Cell Selection (MSCS) scheme.
In MSCS, the RSS threshold for femtocell association, which is referred to as $P_{cs}$, is determined,
and an MS measures the average RSS values of transmitted signals from the neighboring mBSs and fBSs.
If the MS finds some fBSs providing the average RSS larger than $P_{cs}$, the MS associates with the best
fBS among them regardless of the RSS values from the mBSs.
Therefore, the target femtocell radius $d_{f}$ is directly related to $P_{cs}$ by
\begin{equation}
\label{eq:P_cs}
P_{cs} = {P_{f} } \left( {Z d_f} \right)^{ - \alpha },
\end{equation}
where ${P_{f} }$ is the fixed transmission power density of the fBS.
We optimize $d_{f}$ instead of $P_{cs}$ in this paper to simplify the presentation.
Then,
we formulate our parameter optimization problem
for load balancing in hybrid access femtocell networks.
The same framework can be applied for the open access femtocell networks
by setting $\beta = 0$.
Deployment of open and hybrid access femtocells
has an advantage that it can improve the performance of the macrocell users as well as the femtocell owners by transferring the traffic load in the congested macrocells to the femtocells.
Though the open access is allowed, the fMSs expect a differentiated experience
when they use their femtocells at home.
Guaranteeing the benefits of fMSs is very important to motivate the consumers
to buy and install the femtocells.
Therefore, we aim at maximizing the average performance of mMSs
by controlling $\rho$, $d_f$, and $\beta$~(hybrid access only),
while guaranteeing the relative benefits of femtocell owners.
Specifically,
it is assumed that fMSs
expect that their average femtocell throughput should be at least $M$ times larger
than the average throughput of mMSs though they allow the open access.
Denoting the average throughput of an fMS, mMS, and oMS by $\overline {T}_f \left( {\rho ,d_f, \beta } \right)$, $\overline {T}_m\left( {\rho ,d_f,\beta} \right)$, and $\overline {T}_o\left( {\rho ,d_f, \beta } \right)$, respectively, our optimization problem is formulated as:
\begin{xalignat}{2}
\label{eq:original_formulation}
\mathop {\max }\limits_{\rho ,d_f, \beta} & ~ \overline {T}_m \left( {\rho ,d_f, \beta} \right) \\
s.t. \; \notag\\
& \overline {T}_f \left( {\rho ,d_f, \beta} \right) \ge M\overline {T}_m \left( {\rho ,d_f, \beta } \right), \label{eq:const1} \\
& \overline {T}_o \left( {\rho ,d_f, \beta} \right) \ge K\overline {T}_m \left( {\rho ,d_f, \beta} \right), \label{eq:const_hh} \\
& D_h \le d_f \le D_{\max }, \label{eq:const2} \\
& 0 \le \rho\le 1, \label{eq:const3} \\
& 0 \le \beta\le 1. \label{eq:const4}
\end{xalignat}
The macrocell users might not want to associate with fBSs if
the performance is degraded by the open access with fBSs. Therefore, the
constraint~(\ref{eq:const_hh}) is additionally introduced to guarantee the minimum performance of oMSs.
Because oMSs are not the subscribed users, $K$ is configured as a value smaller than $M$, and
we basically assume that $K=1$.
Our optimization framework is a long-term parameter optimization
based on the average system-wide performance metric.
Though a short-term local resource allocation scheme might
improve the performance of the local system efficiently,
our long-term optimization is very meaningful
in the following aspects.
\begin{figure}
\begin{center}
\includegraphics[width=0.35\textwidth]{./SON.eps}
\caption{SON architecture of 3GPP LTE system.}
\label{fig:son}
\end{center}
\end{figure}
First, the parameter optimization involving both the
macrocells and femtocells is generally performed
at a long-term interval due to the system
architecture of the two-tier cellular networks.
In 3GPP LTE system, automatic parameter configuration and optimization
are conducted based on Self Organizing Network~(SON) procedures.
Fig.~\ref{fig:son} illustrates the
SON architecture of 3GPP LTE system,
where macrocell and femtocell base stations
are referred to as Home eNodeB~(HeNB) and eNodeB~(eNB), respectively~\cite{book12Hamalainen}.
SON algorithms can be implemented either
in the end devices, i.e., HeNBs or eNBs, and/or Device Management
and/or Home eNodeB Management System
, and/or Network Management~(NM). However, it is natural that the joint optimization of
macrocell and femtocell parameters is performed by NM in a centralized manner,
because no direct interface between eNB and HeNB exists.
Short-term parameter optimization in the centralized entity, e.g., NM,
is almost impossible due to the limited processing power and
the limited Operations, Administration and Maintenance
bandwidth on the interfaces among the entities.
Consequently, the joint parameter optimization involving both
macrocells and femtocells need to be performed
in the centralized entity at a long-term interval.
Second, short-term resource allocation schemes cannot generally
consider the network-wide performance
due to the excessive overhead and lack of information.
On the other hand, long-term optimization can consider
the network-wide performance thanks to the relatively small overhead
per unit time and relaxed time constraint.
Therefore, our long-term optimization is not
incompatible with but complementary to the short-term
resource allocation schemes
in the sense that the long-term optimization framework can provide good guidelines
for the parameter configuration considering the system-wide average
performance.
Finally, our long-term parameter selection schemes have
the strength in the sense that they can be utilized
in both the self-configuration and/or self-optimization phases.
Note that SON algorithms can be categorized into self-configuration and self-optimization~\cite{3gppTS36.3.9}
according to the functionality and phases.
Self-configuration presents pre-operational procedures including
the initial parameter selection.
Though an initial parameter configuration is essential, an initial parameter
selection based on the local instantaneous optimization is not generally recommended
because the information is very insufficient and unreliable.
On the other hand, our algorithm which does not require the instantaneous local information can be used
for initial parameter selection. The parameters
determined by self-configuration may be updated by the localized self-optimization algorithms in the operational phase.
\section{Numerical Analysis of Two-Tier Cellular Networks}
\label{sec:analysis}
\subsection{Average Throughput of fMS}
\label{sec:femto_model}
In this section, we analyze the average throughput performance of fMSs.
Let us consider an fMS who is located at $r_f$ away from its serving fBS.
As explained in Section~\ref{sec:system_model}, we simply assume that the distribution of fBSs
follows pure SPPP in numerical analysis.
Due to the characteristics of homogeneous Poisson point process~\cite{book93kingman},
the interference measured by a typical fMS is representative of the interference
seen by all other fMSs.
Then, similarly to the SINR models in~\cite{tit0baccelli,pimrc10cheng},
the complementary cumulative distribution function~(CCDF) of an fMS's SINR is given by
\begin{xalignat}{2}
\label{eq:sinr_femto}
F_f \left( {\Gamma |r_f} \right) \buildrel \Delta \over = &~ \Pr \left[ { SINR \ge \Gamma |r_f} \right] \notag\\
= &~ \exp \left( { - sP_N } \right)\exp \left( { - \frac{{2\pi ^2 \lambda _f Z_5^{ - 2} \left( {sP_f } \right)^{2/\alpha _5 } }}{{\alpha _5 \sin \left( {2\pi /\alpha _5 } \right)}}} \right),
\end{xalignat}
where $s = \Gamma \left( {Z_2 r_f } \right)^{\alpha _2 } P_f^{ - 1}$.
The pathloss parameters in the above equation, i.e., $Z_2$, $\alpha_2$, $Z_5$, and $\alpha_5$, are properly chosen
from Table~\ref{table:pathloss} by considering that the fMSs are always located inside the buildings in our system model.
\begin{table}
\centering
\caption{Pathloss parameters }\label{table:pathloss}
\begin{tabular}{|c||c||c|}
\hline
Environment & Exponent & Fixed loss (dB) \\
\hline\hline
Outdoor & $\alpha_1=4$ & $Z_1^{\alpha_1} = 30\log _{10} f_c - 71$\\ \hline
Indoor & $\alpha_2=3$ & $Z_2^{\alpha_2}= 37$\\ \hline
Outdoor-to-indoor & $\alpha_3=4$ & $Z_3^{\alpha_3}= 30\log _{10} f_c - 71 + WL$\\ \hline
Indoor-to-outdoor & $\alpha_4=4$ & $Z_4^{\alpha_4}= 30\log _{10} f_c - 71 + WL$\\ \hline
Indoor-to-indoor & $\alpha_5=4$ & $Z_5^{\alpha_5}= 30\log _{10} f_c - 71 + 2WL$\\ \hline
\end{tabular}
\end{table}
The detailed derivation for (\ref{eq:sinr_femto}) is given in Appendix~\ref{appendix:fMS}.
The probability density of the distance between an fMS and its serving fBS is $r_f$ is given by
$\frac{{2r_f}}{{D_h^2 }}$.
Hence, the average spectral efficiency of an fMS, denoted by $\overline{B}_f$, is calculated as
\begin{equation}
\label{eq:B_f_2}
\overline {B}_f =\sum\limits_{l = 1}^L {\int_0^{D_h } {b_l \left[ {F_{f } \left( {\Gamma _l |r_f} \right) - F_{f } \left( {\Gamma _{l + 1} |r_f} \right)} \right]\frac{{2r_f}}{{D_h^2 }}dr_f} },
\end{equation}
where $b_l$, $\Gamma _{l}$, and L are the spectral efficiency of rate index $l$, the SINR threshold
to utilize the rate index, and the number of available rate sets, respectively.
At a given average spectral efficiency value,
the average throughput of an fMS is degraded when the femtocell resources
are shared with other macrocell users
i.e., oMSs,
in the hybrid (and~open) access mode.
Let us denote the random variable for the number of oMSs who are associated
with an fBS as $N_{o}$.
Then, the average throughput $\overline {T}_f $ is given by
\begin{xalignat}{2}
\label{eq:T_f_1}
\overline {T}_f \left( {\rho ,d_f, \beta } \right)
=& E\left[ {\beta \rho W B_f} \right]
+ E\left[ {\frac{{(1-\beta) \rho W B_f}}{{N_{o} + 1}}} \right] \notag\\
=& \beta\rho W \overline B _f + \left( {1 - \beta } \right)\rho W \overline B_f
E\left[ \frac{1}{N_o + 1} \right],
\end{xalignat}
\noindent where $W$ is the system bandwidth and $B_f$ is the spectral efficiency of an fMS.
In Eq.~(\ref{eq:T_f_1}), the dedicated resources to fMS contribute to the first term while the second term is due to the shared resources between fMS and oMS.
In addition, the equality holds because the spectral efficiency of an fMS, i.e., $\overline B_f$, is independent of the
number of $N_{o}$.
If we refer to the service area of a given femtocell as $y$,
the number of oMSs in the femtocell's service area
follows Poisson random variable with the mean $\lambda _u y$, and
the probability mass function~(pmf) of $N_{o}\left(y\right)$ is given by
\begin{equation}
\label{eq:macro_dist}
f_{N_{o}\left(y\right)}\left[ {k} \right] \sim \frac{{\left( { \lambda _uy } \right)^k {\mathop{\rm e}\nolimits} ^{ - \lambda _uy } }}{{k!}},
\end{equation}
and we obtain the expectation value as
\begin{equation}
\label{eq:N_derive}
E\left[ { \left. {\frac{1}{{N_{o}\left(y\right) + 1}}} \right| y } \right] = \sum\limits_{k = 0}^\infty { \frac{{\left( {\lambda _u y} \right)^k e^{ - \lambda _u y} }}{{\left(k+1\right)k!}} = \frac{{1 - e^{ - \lambda _u y} }}{{\lambda _u y}}}.
\end{equation}
Let us denote the average service area of a femtocell as $x = E\left[ y \right]$.
From~(\ref{eq:T_f_1}) and~(\ref{eq:N_derive}),
we obtain the approximated value for the average throughput of an fMS as follows:
\begin{xalignat}{2}
\label{eq:T_f_2}
\overline T _f \left( {\rho ,d_f, \beta } \right)
= &~ \beta\rho W \overline B _f + \left( {1 - \beta } \right)\rho W \overline B _f E_y \left[ {\frac{{1 - e^{ - \lambda _u y} }}{{\lambda _u y}}} \right] \notag\\
\cong &~ \beta \rho W \overline B _f + \frac{{\left( {1 - \beta } \right)\rho W \overline B _f \left( {1 - e^{ - \lambda _u x} } \right)}}{{\lambda _u x}}.
\end{xalignat}
The average throughput of an fMS in open access femtocell
is obtained by setting $\beta=0$.
Here, we obtain the average service area, i.e., $x$, which is a function of the target
service radius $d_f$.
A macro user associates with an mBS only when
no fBS exists within $d_f$ meters from the user.
According to the property of SPPP, the probability that
the distance $R$ between a specific mMS and its nearest fBS is larger than $r_1$ is given by
\begin{equation}
\label{eq:nearest_cdf}
\Pr \left( {R > r_1 } \right) = \Pr \left( {{\rm{No~fBS~closer~than~}}r_1 } \right) = e^{ - \pi \lambda_f r_1^2 }.
\end{equation}
Therefore, the probability that a macrocell user is covered by an fBS is given by
${1 - e^{ - \pi d_f^2 \lambda _f } }$, and
it is equivalent to the fact that the fBSs cover the area of $A_m\left( {1 - e^{ - \pi d_f^2 \lambda _f } } \right)$ on average.
Because the average number of fBSs in a macrocell area is $A_m \lambda _f$,
the average area of a femtocell region can be approximated by
\begin{equation}
\label{eq:x}
x\left( d_f \right) = \frac{{1 - e^{ - \pi d_f^2 \lambda _f } }}{{\lambda _f }}.
\end{equation}
Because $d_f$ and $x$ have an one-to-one relationship, we use $x$
as the control parameter instead of $d_f$ in the rest of the paper for the simplicity
of presentation.
\subsection{Average Throughput of mMS}
\label{sec:macro_model}
In this section, we model the average throughput performance of an mMS,
where mMS is defined by the macrocell user who is currently associated
with the mBS.
Let us refer to the distance between an mMS and its serving mBS
as $r_m$.
From the assumption that each fBS fully covers its indoor home area,
SINR CCDF of an
mMS for a given $r_m$ is obtained by
\begin{equation}
\label{eq:F_m}
F_{m } \left( {\Gamma |r_m} \right) = \exp \left( { - \frac{{\Gamma \left( {P_N } \right)}}{{P_m \left( {Z_1 r_m} \right)^{ - \alpha_1 } }}} \right).
\end{equation}
As shown in Table~\ref{table:pathloss}, $Z_1$ and $\alpha_1$ are the fixed pathloss value and the pathloss exponent in outdoor environments,
respectively. The detailed derivation for the above equation is shown in Appendix~\ref{appendix:mMS}.
The probability that a macrocell user becomes an mMS, which
is referred to as $p_{mMS}$, is given by
\begin{equation}
\label{eq:p_mms}
p_{mMS} = e^{ - \pi \lambda _f d_f^2 } = 1 - \lambda _f x,
\end{equation}
where $x$ is the average service area of a femtocell derived in (\ref{eq:x}).
As shown in (\ref{eq:p_mms}) that $p_{mMS}$ is independent of the location of
the macrocell user, i.e., the distance between the user and the mBS,
due to the random distribution of fBSs.
Therefore, if we refer to the distance between an mMS and its serving mBS as $r_m$,
the probability density function~(PDF) of $r_m$ is also given by
\begin{equation}
\label{eq:f_R_m}
f_{R_m } \left( {r_m } \right) = \frac{{2r_m }}{{D_m^2 }}.
\end{equation}
From (\ref{eq:F_m}) and (\ref{eq:f_R_m}), the average spectral efficiency of an mMS is given by
\begin{equation}
\label{eq:B_m}
\overline {B}_m =\sum\limits_{l = 1}^L {\int_0^{D_m } {b_l \left[ {F_{m } \left( {\Gamma _l |r_m} \right) - F_{m } \left( {\Gamma _{l + 1} |r_m} \right)} \right]\frac{{2r_m}}{{D_m^2 }}dr_m} },
\end{equation}
where $\Gamma _{L + 1}=\infty$.
In order to calculate the above equation, we calculate
\begin{xalignat}{2}
\label{eq:integration}
\int_0^{D_m } {F_{m } \left( {\Gamma_l |r_m} \right)r_m} dr_m
& = \int_0^{D_m } {r_m e^{ - \beta_{l} r_m^\alpha}} dr_m,
\end{xalignat}
where $\beta_l$ is defined by $P_N\Gamma_l Z_{1}^{\alpha_1} P_m^{ - 1}$.
By substituting $- \beta_{l} r^{\alpha_1}$ with $y$, the above equation is obtained by
\begin{xalignat}{2}
\label{eq:integration2}
\int_0^{\beta _l D_m^{\alpha_1} } {\frac{1}{{\alpha_1 \beta _l }}\left( {\frac{y}{{\beta _l }}} \right)^{\frac{{2 - \alpha_1 }}{\alpha_1 }} e^{-y}} dy &= \frac{ \beta_{l} ^{ - \frac{2}{\alpha_1 }} }{\alpha_1} G \left( {\frac{2}{\alpha_1 },\beta_l D_m^{\alpha_1} } \right),
\end{xalignat}
where $G \left( {a, b} \right) \buildrel \Delta \over = \int_0^b {t^{a - 1} e^{-t}dt}$ is the incomplete gamma function.
Although femtocell deployment does not change the average spectral efficiency of an mMS,
the average throughput of an mMS can be improved by deploying the femtocells, because
the number of mMSs sharing the macrocell resource is reduced by relocating some macrocell users, i.e., oMSs,
to the femtocells.
In our system model, the average number of macrocell users is given by
$\overline N_u = A_m \lambda _u$.
Let us refer to
the number of mMSs in a macrocell area for a given $x$ as
$N_m$.
From (\ref{eq:p_mms}),
the expectation of $N_m$ is given by $\overline N_m \left(x\right) = A_m \lambda _u \left( {1 - \lambda _f x} \right)$.
We approximate $N_m$ as a Poisson random variable with the mean of $\overline N_m$, and
hence, the average throughput of an mMS is given by
\begin{xalignat}{2}
\label{eq:T_m}
\overline T _m \left( {\rho ,x} \right)
& = \sum\limits_{n = 1}^\infty {\frac{{\left( {1 - \rho } \right) W \overline B _m }}{n}} \frac{{nf_{N_m } \left[ n \right]}}{{\sum\limits_{k = 1}^\infty {kf_{N_m } \left[ k \right]} }} \notag\\
& \cong \frac{{\left( {1 - \rho } \right) W \overline B _m \left( {1 - e^{ - A_m \lambda _u \left( {1 - \lambda _f x} \right)} } \right)}}{{A_m \lambda _u \left( {1 - \lambda _f x} \right)}}.
\end{xalignat}
\subsection{Average Throughput of oMS}
\label{sec:oa_model}
In order to complete the load balancing problem formulated in~(\ref{eq:original_formulation}),
we analyze the average throughput of an oMSs and the maximum service radius of femtocells, i.e.,
$\overline T_{o}$ and $D_{\max}$, respectively.
When the target femtocell service radius is configured as $d_f$,
a macrocell user associates with its nearest fBS if the
distance between the user and fBS is equal to or smaller than $d_f$.
From (\ref{eq:p_mms}), the probability that a
macrocell user becomes an oMS is given by $\lambda_fx$.
Let us refer to the distance between an oMS and its serving fBS
as $r_o$.
Then, the conditional probability density function of $r_{o}$
is given by
\begin{equation}
\label{eq:f_R_oa}
f_{R_{o} } \left( {r_{o} |x} \right) = \left\{ {\begin{array}{*{20}c}
{\frac{{2\pi \lambda _f r_{o} e^{ - \pi \lambda _f r_{o}^2 } }}{{\lambda _f x}},} & {0 \le r_{o} \le \sqrt { \frac{{\ln \left( {1 - \lambda _f x} \right)}}{{- \pi \lambda _f }}}}, \\
{0,} & \mathrm{otherwise}. \\
\end{array}} \right.
\end{equation}
When $r_{o}$ is given,
the SINR CCDF of an oMS is given by:
\begin{equation}
\label{eq:sinr_femto_OA_1}
\begin{array}{l}
F_o \left( {\Gamma \left| {r_o } \right.} \right)
\\ = \exp \left( { - \frac{{\pi \lambda _f \sqrt {sP_f } }}{{Z_I^2 }}\left( {\frac{\pi }{2} - \tan ^{ - 1} \left( {\frac{{Z_I^2 r_o^2 }}{{\sqrt {sP_f } }}} \right)} \right) - sP_N } \right), \\
\end{array}
\end{equation}
where
\begin{equation}
\left( {Z_I ,s} \right) = \left\{ {\begin{array}{*{20}c}
{\left( {Z_4 ,P_f^{ - 1} \Gamma \left( {Z_4 r_o } \right)^{\alpha _4 } } \right),} & {r_o \ge D_h, } \\
{\left( {Z_5 ,P_f^{ - 1} \Gamma \left( {Z_2 r_o } \right)^{\alpha _2 } } \right),} & \mathrm{otherwise.} \\
\end{array}} \right.
\end{equation}
The detailed derivation for the above equation is found in Appendix~\ref{appendix:oaMS}.
Then, we obtain the average spectral efficiency of an oMS.
If we refer to the probability that an oMS chooses the rate index $l$
as $f_L \left[ {l|r_{o} } \right] = F_{o} \left( {\Gamma_l |r_{o} } \right) - F_{o} \left( {\Gamma_{l + 1} |r_{o} } \right)$,
the average spectral efficiency of an oMS is given by
\begin{equation}
\label{eq:B_o_done}
\overline B _{o} \left( x \right) = \sum\limits_{l = 1}^L {\int\limits_0^{d_f\left( x \right) } {b_l f_L \left[ {l|r_{o} } \right]f_{R_{o} } \left( {r_{o} | x} \right)dr_{o} } },
\end{equation}
where $f_{R_{o} } \left( {r_{o} | x} \right)$ is shown in (\ref{eq:f_R_oa}).
We refer to the random variable representing the number of oMSs in a
femtocell area as $N_{o}$.
As done for the average throughput analysis of an fMS
in (\ref{eq:T_f_2}), we approximately assume that $N_{o}$ is
a Poisson random variable with the mean of $\lambda_ux$.
Because $N_{o} + 1$ users including an fMS equally share
the femtocell resources in the open access mode
and only $1-\beta$ fraction of intra-femtocell resources are allowed
to the oMSs,
the average throughput of oMSs is expressed by
\begin{equation}
\label{eq:th_oa_0}
\overline T _{o} \left( {\rho ,x,\beta } \right) = \sum\limits_{n_{o} = 1}^\infty {\frac{{\rho \left( {1 - \beta } \right) W \overline B _{o} \left( x \right)}}{{n_{o} + 1}}} \frac{{n_{o}f_{N_{o} } \left[ n_{o} \right]}}{{\sum\limits_{k = 1}^\infty {kf_{N_{o} } \left[ k \right]} }},
\end{equation}
where
{\small
\begin{equation}
\label{eq:th_oa_1}
\sum\limits_{n = 1}^\infty {\frac{{nf_{N_o } \left[ n \right]}}{{n + 1}}} = \frac{1}{{\overline N _o }}\sum\limits_{n = 2}^\infty {\left( {n - 1} \right)} \frac{{\overline N _o ^n e^{ - \overline N _o } }}{{n!}} = \frac{{\lambda _u x + e^{ - \lambda _u x} - 1}}{{\lambda _u x}}.
\end{equation}}
From (\ref{eq:th_oa_0}) and (\ref{eq:th_oa_1}), $\overline T _{o}$ is obtained by
\begin{equation}
\label{eq:th_oa}
\overline T _{o} \left( {\rho ,x,\beta } \right) = \frac{{ \left( 1-\beta \right) \rho W \overline B _{o} \left( x \right)\left( {\lambda _u x + e^{ - \lambda _u x} - 1} \right)}}{{\left( {\lambda _u x} \right)^2 }}.
\end{equation}
Furthermore, we analyze the maximum service radius of a femtocell, i.e., $D_{max}$,
to complete the problem formulation in (\ref{eq:original_formulation}).
$D_{max}$ is an important parameter which determines the range of our design
parameter $d_f$ (or $x$).
As described in Section~\ref{sec:modeling}, we define $D_{max}$ as the maximum
target service radius where the average outage rate of an oMS is less than or equal to
$O_{\max}$.
We assume that an outage occurs when the
instantaneous SINR is less than the threshold for the lowest transmission rate,
i.e., $\Gamma_1$ in Table~\ref{table:rate_sets}.
From the definition, average outage rate of an oMS with the given target service radius $d_f$ is calculated by
\begin{equation}
\label{eq:D_max_1}
\overline O_{o} \left( d_f \right) = \int\limits_0^{d_f } {\left( {1 - F_o \left( {\Gamma _1 \left| {r_o } \right.} \right)} \right)f_{R_o } \left( {r_o \left| d_f \right.} \right)} dr_o,
\end{equation}
which looks very similar to~(\ref{eq:B_o_done}).
$D_{\max}$ is obtained from the equation $\overline O _o \left( D_{\max} \right) = O_{\max }$.
Though $D_{\max}$ cannot be given in a closed form, the near-optimal solution for~(\ref{eq:D_max_1}) can
easily be obtained by using the simple binary search algorithm because
$\overline O _o \left( d_f \right)$ is a monotonically increasing function of $d_f$.
Because any mobile stations which are in the femtocell coverage should attached to femtocell, SINR of the oMS is likely to worse as the service radius of a femtocell, $d_f$, is larger when it is larger than threshold, i.e., the radius that a reference signal received power from the nearest fBS and mBS are about the same.
The detailed description for the binary search algorithm is omitted due to the space limitation.
\section{Optimization in Open Access Femtocell Networks}
\label{sec:optimization_open}
\subsection{Optimization of Parameters in Open Access Femtocell Networks}
\label{subsec:optimization}
In this section, we prove that our optimization problem in open access mode becomes
a single variable convex problem in some typical environments, and the optimal parameters are obtained.
In this section, $\beta=0$ because we consider the open access femtocells.
If we define $\overline {t}_m \left( {x } \right)$ and $\overline {t}_{fo} \left( {x } \right)$ as
\begin{equation}
\label{eq:small_th_m}
\overline {t}_m \left( x \right) = \frac{{\overline T _m \left( {\rho ,x,\beta=0} \right)}} {W\left({1 - \rho }\right)}
\end{equation}
and
\begin{equation}
\label{eq:small_th_fo}
\overline {t}_{fo} \left( x \right) = \min \left( {\frac{{\overline {T} _f \left( {\rho ,x,\beta=0} \right)}}{M W \rho },\frac{{\overline T _o \left( {\rho ,x,\beta=0} \right)}}{K W \rho }} \right)
\end{equation}
the optimization problem~(\ref{eq:original_formulation}) is rephrased by
\begin{xalignat}{2}
\label{eq:revised_formulation}
\mathop {\max }\limits_{\left( {\rho ,x} \right)} & ~ \left( 1-\rho \right) \overline {t}_m \left( {x } \right) \\
s.t. \; \notag\\
& \rho \overline {t}_{fo} \left( {x } \right) \ge \left( 1-\rho \right) \overline{t}_m \left( {x } \right), \label{eq:const1_r} \\
& X_{\min} \le x \le X_{\max}, \label{eq:const2_r} \\
& 0 \le \rho\le 1,
\label{eq:const3_r}
\end{xalignat}
where $X_{\min} = x\left( {D_h } \right)$ and $X_{\max} = x\left( {D_{\max} } \right)$ from (\ref{eq:x}).
\newtheorem{prop}{Proposition}
\begin{prop}\label{prop:optimal_bw}
The optimal $\rho^{*}$ which maximizes the objective in~(\ref{eq:revised_formulation}) is given by
$\rho^* = \frac{{ \overline{t}_m \left( {x^* } \right) }}{{ \overline{t}_{fo} \left( {x^* } \right) + \overline{t}_m \left( {x^* } \right) }}$,
where $x^{*}$ is the optimal value of $x$.
\end{prop}
\begin{IEEEproof}
See Appendix~\ref{appendix:proof1}.
\end{IEEEproof}
As inferred by (\ref{eq:small_th_fo}) and (\ref{eq:const1_r}), the performance of our load balancing
algorithm is always limited by the throughput requirement of an fMS if
$K\overline T _f \left( {\rho ,x, 1 } \right) \le M\overline T _o \left( {\rho ,x, 1} \right)$
for all $x \in \left[ { X_{\min}, X_{\max} } \right]$.
We define such cases by the fMS's requirement-limited environments.
Then, the following proposition holds.
\begin{prop} \label{prop:convex}
In the fMS's requirement-limited environments, the optimization
problem~(\ref{eq:revised_formulation}) is a convex optimization problem.
\end{prop}
\begin{IEEEproof} See Appendix~\ref{appendix:proof2}.
\end{IEEEproof}
From the above proposition, the optimal solution of $x$ can be efficiently obtained by using the
standard methods used for solving the convex optimization~\cite{book04boyd} if
the given environment is the fMS's requirement-limited environment.
We expect that the two-tier cellular networks mostly operate in the fMS's requirement-limited
environments, because the oMS's throughput requirement is easily satisfied with the
reasonably large fMS's benefit requirement, i.e., $M$.
However, if the given environment is not the fMS's requirement-limited environment,
we should obtain the near-optimal solution of $x$ using the inefficient exhaustive search algorithm.
To check whether a given environment is the fMS's requirement-limited environment,
the following proposition could be useful.
\begin{prop} \label{prop:convex_condition}
If we define $C\left( x \right) \buildrel \Delta \over = \frac{{ \left( {1 - e^{ - \lambda _u x} } \right)}}{{\lambda _u x}}$ and
$D\left( x \right) \buildrel \Delta \over = \frac{{ \left( {\lambda _u x + e^{ - \lambda _u x} - 1} \right)}}{{\left( {\lambda _u x} \right)^2 }}$, $\frac{{\overline B _f C\left( {X_{\min}} \right)}}{{\overline B _o \left( {X_{\max}} \right)D\left( { X_{\min}} \right)}} \le \frac{M}{K}$
is the sufficient condition that a given environment is the fMS's requirement-limited environment.
\end{prop}
\begin{IEEEproof}
See Appendix~\ref{appendix:proof3}.
\end{IEEEproof}
From the (near) optimal value $x^*$, the service area control is implemented by
setting $P_{cs}$ according to~(\ref{eq:P_cs}) and~(\ref{eq:x}).
The optimal $\rho^*$ is configured according to Proposition~\ref{prop:optimal_bw}.
The optimal solution $x^*$ has the following property.
\begin{prop} \label{prop:D_max_optimal}
$\frac{{\overline{N}_{f} \overline {B}_f }}{{M\overline {B}_m }} > 1$, where $\overline{N}_{f}$ is the average number of fBSs in a macrocell area, is a sufficient condition that the optimal solution of the problem~(\ref{eq:revised_formulation}) is given by
$x^* = X_{\max}$ in the typical fMS's requirement limited environments where
the average number of users in a macrocell is larger than that in a single femtocell.
\end{prop}
\begin{IEEEproof}
See Appendix~\ref{appendix:proof4}.
\end{IEEEproof}
Proposition~\ref{prop:D_max_optimal} can be interpreted as follows.
The expansion of the femtocell service area is encouraged if the average spectral efficiency of
an fMS is much larger than that of an mMS,
and the expansion of the femtocell service area is also encouraged if there are many femtocells in
the system because the sum of performance gain in the system is approximately proportional to the number of femtocells.
On the other hand, large benefit requirements from
fMSs, i.e., $M$, can limit the service area expansion to guarantee the performance of femtocell owners who are the premium users.
\subsection{Extension Considering Interference Thinning (Thin)}
\label{sec:interference_management}
As described in Proposition~\ref{prop:D_max_optimal}, the optimal performance of our load balancing
scheme is determined by the physical limitation $D_{\max}$ in many cases.
Therefore, we expect that the efficiency of our load balancing scheme can be enhanced
if the maximum coverage of femtocell is increased by applying an interference thinning scheme,
which reduces the interference by limiting the resource usage in fBSs while all the BSs fully utilize the given resources in our basic model.
In our interference thinning scheme,
the probability that an fBS uses a specific resource block is limited to $\theta$, to reduce the interference among femtocells.
The channel condition is improved by using $\theta < 1$, but using small $\theta$ can reduce the aggregate throughput by decreasing chances for packet transmissions.
Our proposed scheme including the interference thinning scheme is referred to as OA-Thin.
The objective and constraints of OA-Thin is the same as those of the original problem formulation in OA scheme,
but we optimize the new control parameter, i.e., $\theta$,
as well as $\rho$ and $d_f$ to maximize our objective while satisfying the constraints.
Some parameters in our analytic model are updated by considering $\theta$.
The SINR distribution of an fMS and an oMS, i.e., $F_f \left( {\Gamma |r_m } \right)$ in (\ref{eq:sinr_femto})
and $F_o \left( {\Gamma |r_o } \right)$ in (\ref{eq:sinr_femto_OA_1}), are updated by replacing
$\lambda_f$ in the equations with $\theta\lambda_f$.
Furthermore, $D_{\max}$ in is also updated considering $\theta$.
It is difficult to show that finding the optimal solution of OA-Thin scheme is a convex problem
in the fMS's requirement-limited environments as described in Proposition~\ref{prop:convex}.
However, if $\theta$ is given, Proposition~\ref{prop:convex} holds, and
we can find the optimal $x^{*}\left( \theta \right)$ and $\rho^{*}\left( \theta \right)$ efficiently
in the typical environments where the operation is dominated by the benefit requirement
of fMSs.
Therefore, we repeatedly solve the convex optimization problems with various candidate $\theta$ values,
and $\theta$ which minimizes the objective function is chosen as the suboptimal $\theta^{*}$.
\subsection{Analysis of Optimal Parameters}
\label{sec:optimal}
\begin{figure*}
\centering \subfigure[Optimal femtocell coverage ($d_f^*$).]{
\includegraphics[width=0.3\textwidth]{./OA_df_solution_corrected.eps}\label{fig:d_f}}\hfill
\subfigure[Optimal resource utilization ratio ($\theta^*$).]{
\includegraphics[width=0.3\textwidth]{./OA_theta_solution_corrected.eps}\label{fig:theta}}\hfill
\subfigure[Optimal amount of resources dedicated to femtocells ($\rho^*$).]{
\includegraphics[width=0.3\textwidth]{./OA_rho_solution_corrected.eps}\label{fig:w_f}}
\caption{Optimal parameters.} \label{fig:optimal_parameters}
\end{figure*}
In this section, we show the optimal parameters of the proposed schemes based on our analytic model.
The basic parameters shown in Table~\ref{table:parameters} are used in the evaluations
unless mentioned otherwise, and we obtain the results with various $M$ values, where
fMSs require $M$ times higher average throughput than the average throughput of mMS.
Fig.~\ref{fig:d_f} shows the optimal service radius $d_f^*$ given by
the optimization of OA and OA-Thin schemes.
The straight lines in the figure represent $D_{\max}$ with the given average number of fBSs which is referred to as $\overline N_f$.
As $\overline N_f$ increases, $D_{\max}$ value decreases due to the increased interference.
For all $\overline N_f$ values, OA scheme determines to use $D_{\max}$ as the service radius of femtocells when $M$ is not very large.
If $M$ exceeds some threshold, the optimal femtocell radius becomes smaller than $D_{\max}$,
because sharing the femtocell resources with many oMSs
is not an efficient method to provide a large relative benefit to the fMSs.
In the sense of $\overline N_f$, $D_{\max}$ is preferred in the wider range of $M$ values when
$\overline N_f$ is large, because the offloading gain of using femtocells is more significant
with the large number of fBSs.
These results are the same results which are inferred by Proposition~\ref{prop:D_max_optimal}.
The optimal resource utilization ratio of femtocells, i.e., $\theta^*$, in OA-Thin scheme
is shown in Fig.~\ref{fig:theta}.
The needs for interference management is larger in the environments where many fBSs exist.
Therefore, Fig.~\ref{fig:d_f} shows that OA-Thin scheme chooses to expending the maximum service radius by applying $\theta < 1$ when there are $30$ or $50$ fBSs in average.
In the mean time, the interference thinning is not effective
when $M$ is very large, because it is difficult to satisfy the requirement of $M$ if fBSs use the partial resources in the femtocells.
Accordingly, Fig.~\ref{fig:theta} shows that the femtocells are required to fully utilize the dedicated femtocell resources when $M$ is large.
The optimal ratio of resources dedicated to femtocells, i.e., $\rho^*$, is shown in
Fig.~\ref{fig:w_f}. For the region where $d_f^*$ is fixed, $\rho^*$ proportionally increases
as $M$ increases. However, in the middle region where $d_f^*$
increases, $\rho^*$ decreases to properly maximize the average throughput of mMSs
while meeting the requirements for the fMS's performance.
\section{Optimization in Hybrid Access Femtocell Networks}
\label{sec:optimization_hybrid}
In hybrid access femtocell networks,
we optimize $\beta$ as well as $x$ and $\rho$.
The analysis results for the open access femtocell networks in
the previous section are used in the optimization procedures
for the hybrid access femtocell networks.
\begin{prop}\label{prop:optimal_hybrid}
When the target femtocell service area $x$ is given, the optimal $\rho^{*}$ and $\beta^{*}$
which maximize the objective in (\ref{eq:original_formulation}) are given as follows:
\begin{equation}
\beta ^* \left( x \right) = \left\{ {\begin{array}{*{20}c}
\frac{{ - C\left( x \right) + D\left( x \right)}}{{B\left( x \right) - C\left( x \right) + D\left( x \right)}},
& D\left( x \right) \ge C\left( x \right), \\
0, & \mathrm{otherwise}, \\
\end{array}} \right.
\end{equation}
and
\begin{equation}
\rho ^* \left( x \right) = \left\{ {\begin{array}{*{20}c}
\frac{{A\left( x \right)}}{{\frac{{B\left( x \right)D\left( x \right)}}{{B\left( x \right) - C\left( x \right) + D\left( x \right)}} + A\left( x \right)}},
& D\left( x \right) \ge C\left( x \right), \\
\frac{{A\left( x \right)}}{{D\left( x \right) + A\left( x \right)}}, & \mathrm{otherwise}, \\
\end{array}} \right.
\end{equation}
where
$A\left( x \right) \buildrel \Delta \over = \frac{{\overline B _m \left( {1 - e^{ - A_m \lambda _u \left( {1 - \lambda _f x} \right)} } \right) }}{{A_m \lambda _u \left( {1 - \lambda _f x} \right)}}$,
$B\left( x \right) \buildrel \Delta \over = \frac{{\overline B _f }}{M}$,
$C\left( x \right) \buildrel \Delta \over = \frac{{\overline B _f \left( {1 - e^{ - \lambda _u x} } \right)}}{{M\lambda _u x}}$, and
$D\left( x \right) \buildrel \Delta \over = \frac{{\overline B _{o} \left( x \right)\left( {\lambda _u x + e^{ - \lambda _u x} - 1} \right)}}{{K\left( {\lambda _u x} \right)^2 }}$.
\end{prop}
\begin{IEEEproof}
See Appendix~\ref{appendix:proof5}.
\end{IEEEproof}
From Proposition~\ref{prop:optimal_hybrid} and the original problem formulation in (\ref{eq:original_formulation}),
the load balancing problem in hybrid access femtocell can be treated as the single
variable optimization with the control parameter $x$.
Unfortunately, the above optimization problem is not a convex optimization problem.
Therefore, we find a near-optimal solution by calculating the objective function over the
feasible region of $x$.
Because the original problem has been simplified to a single variable problem and
the feasible region of $x$ is bounded,
we can find the near-optimal solution $x^*$
without excessively complex computations.
\begin{figure}
\begin{center}
\includegraphics[width=0.3\textwidth]{./HA_M_K_result_M_corrected.eps}
\caption{Ratio between the average throughputs of oMS and mMS.}
\label{fig:Ratio_MK}
\end{center}
\end{figure}
The performance gain of mMS and fMS in the hybrid access femtocell networks
is achieved at the cost of the average performance degradation of an oMS
by using $\beta>0$.
Fig.~\ref{fig:Ratio_MK} describes the ratio between the throughputs of
oMSs and mMSs
in open access femtocells and hybrid access femtocells based on analysis.
oMSs in the open access femtocells enjoy much larger throughput performance than oMSs in the hybrid access femtocells
thanks to the resource-fair intra scheduling of the fBS.
However, it is unfair that the oMSs achieve such large throughput because
the oMSs and mMSs are actually the same type of users who do not pay any cost for the femtocell deployment.
On the other hand, the hybrid access fBS distinguishes the fMS from the oMSs and the average throughput of
oMSs are properly controlled so that the similar quality of services are provided to the mMSs and oMSs.
Fig.~\ref{fig:Ratio_MK} shows that the average throughput of an oMS in the hybrid access femtocell networks is
exactly $K$ times larger than that of an mMS, where $K$ is generally a small value.
\section{Performance Evaluation}
\label{sec:evaluation}
\subsection{Evaluation Environments and Comparing Schemes}
\label{sec:eval_environments}
In this section, we evaluate our proposed schemes based on both numerical analysis and computer simulations.
As described in Section~\ref{sec:system_model}, macrocell users and fBSs are randomly deployed according
to SPPP, while one constraint that the distance between the fBSs should be less than $2D_h$
is additionally given in the simulation settings.
The channel model used for the numerical analysis and simulation is described in Section~\ref{sec:system_model}, and the basic values of the evaluation parameters are provided in Table~\ref{table:parameters}~\cite{twc10jo,itur.m.1225}.
We consider the random mobility of macrocell users in the simulations, and the performances of users located in the interested area, i.e., a single macrocell area, are considered for the performance analysis.
In our evaluations, the proposed schemes are compared with some comparing schemes.
The basic comparing scheme is CoRSSI where the mBSs and fBSs share the
same bandwidth, and a user associates with the cell which provides the best signal strength
including both macrocells and femtocells.
CoRSSI is excellent in the aspect of sum capacity by fully reusing the
bandwidth, but the benefit achieved by
mMS can be smaller than the dedicated bandwidth based schemes
due to the limited offloading gain as will be shown in
Section~\ref{sec:simulations}.
In order to show the maximum offloading gain in a co-channel deployment, we also introduce CoLB~(Cochannel Load Balancing) scheme where
the system promotes the users to access femtocell as much as possible
while the basic requirement for service coverage
of femtocell is satisfied. The details of CoLB is described in Section~\ref{sec:simulations}.
DivRSSI is the scheme which assigns the dedicated orthogonal resources to femtocells like the proposed schemes, and each MS associates with the BS which provides the best RSSI value like CoRSSI. In DivRSSI,
the bandwidth is divided into the macro and femtocell resources with
the optimal ratio, i.e., $\rho^*$, while maintaining the constraints $\overline T_f \ge M \overline T_m$ and $\overline T_f \ge K \overline T_o$.
The optimal $\rho^*$ is chosen based on the simulation results
in DivRSSI because no numerical analysis model exists for DivRSSI.
Finally, femtocells do not allow any open or hybrid access of oMSs
to access femtocells in CoCA and DivCA, where CA stands for Closed Access.
Similarly to DivRSSI, the optimal bandwidth dedicated to femtocell is applied
for DivCA based on the simulation results, while the whole
resources are shared by macrocell and femtocell users in CoCA.
\subsection{Analysis and Simulation Results}
\label{sec:simulations}
\begin{figure*}
\centering \subfigure[Error between analysis and simulation results.]{
\includegraphics[width=0.3\textwidth]{./validation.eps}\label{fig:validation}}\hfill
\subfigure[Average throughput of mMS vs. required fMS's benefit, i.e., M.]{
\includegraphics[width=0.3\textwidth]{./updated_HA.eps}\label{fig:TH_m}}\hfill
\subfigure[Average throughput of mMS vs. number of fBSs, i.e., $N_f$.]{
\includegraphics[width=0.3\textwidth]{./OA_TH_m_result_Nf_10.eps}\label{fig:TH_Nf}}
\caption{Basic simulation results.} \label{fig:basic_simulations}
\end{figure*}
In this section, we evaluate the performance of proposed schemes using numerical
analysis.
The default evaluation parameters specified in Table~\ref{table:parameters}
are used unless mentioned otherwise.
First, we show that the
numerical analysis is valid in spite of the simplifying
assumptions and approximations in the numerical analysis.
We obtain the average spectral efficiency and average throughput of an fMS,
oMS, and mMS when the service radius $d_f$ values are given.
After simulations with $10,000$ distributions, we compare the average simulation
results with the analysis result as shown in Fig.~\ref{fig:validation}.
The validation results indicate that
the errors between the results of the numerical analysis and computer simulation
are negligible, i.e., less than $1~\%$.
Fig.~\ref{fig:TH_m} shows the average throughput of an mMS.
We find that the performance of mMSs is enhanced by utilizing the proposed
schemes in most regions. When the required benefit from fMSs is small,
the more aggressive traffic offloading from macrocell users is feasible.
Therefore, the performance gains of OA and OA-Thin are more significant when $M$ is small.
Among the comparing schemes CoRSSI provides
comparable or even better average throughput performance to mMSs than the proposed schemes when the required fMS's benefit exceeds a certain value.
This result shows that co-channel deployment could be more efficient
to guarantee very large performance benefits to fMSs.
However, this phenomenon happens when $M$ is very large, e.g., $M > 50$ in this example, and such large benefits for fMSs might not be required in reality.
HA(-Thin) schemes achieve the better average throughput of an mMS than OA(-Thin) schemes.
Especially, the gain of hybrid access femtocell is still very significant
even in the environments where $M$ is very large, while the open access femtocell's offloading gain is limited in the environments.
In HA(-Thin) schemes, the performance of mMSs and fMSs are
improved by preventing the oMSs from achieving the performance gain
much more than necessary.
OA-Thin and HA-Thin schemes obtain the further performance gain by
improving the SINR status and extending the maximum femtocell coverage.
As discussed in Section~\ref{sec:optimal}, the partial utilization of femtocell resources, i.e., OA/HA-Thin, is preferred when $M$ is small.
The impact of the number of fBSs in a macrocell area is shown in Fig.~\ref{fig:TH_Nf}.
In the open access femtocell networks, the performance of
mMS is enhanced as the number of fBSs increases because the chances for
the offloading of macrocell's traffic increases.
As the previous propositions
implicate, the interference thinning in OA/HA-Thin becomes useful
when the number of fBSs is
large where the load balancing efficiency is maximized.
Obviously, the performance of CoCA scheme, i.e., co-channel deployment based on the
closed access femtocells, is rapidly degraded as $N_f$ increases due to the excessive
co-channel interference.
Load balancing gain can also be enhanced
in co-channel deployment scenario by expanding
the service coverage of femtocells.
CoLB scheme expands the service coverage of femtocells
as much as possible while the basic requirement for
signal quality is satisfied.
A weight factor for femtocell access, i.e., $\delta_f$,
is notified by the system under CoLB scheme, and a macrocell user associates
with a femtocell if $\delta_f RSSI_{f,\max} > RSSI_{m}$, where
$RSSI_{f,\max}$ is the maximum received signal strength from
the nearby fBSs and $RSSI_{m}$ is the received signal strength from the mBS, respectively.
CoLB is identical to CoRSSI when $\delta_f = 1$.
Though the offloading gain obtained
in the macrocell tier increases as $\delta_f$ increases, the maximum $\delta _f$ is limited by the service quality constraint, e.g., average outage rate, of oMSs.
For a fair comparison with the proposed schemes, we use
the maximum $\delta _f$ which satisfies $\overline O _o \left( {\delta _f } \right) \le O_{\max }$, where the outage rate threshold, i.e., $O_{\max }$, is the same as that used in the proposed schemes to limit the maximum femtocell service radius, i.e., $D_{\max}$.
\begin{figure}
\centering \subfigure[Average outage rate an oMS in CoLB scheme.]{
\includegraphics[width=0.3\textwidth]{./Out_CoLB_TH.eps}\label{fig:out_colb}}
\subfigure[Average throughput of mMS using CoLB scheme.]{
\includegraphics[width=0.3\textwidth]{./CoLB_TH.eps}\label{fig:th_colb}}
\caption{Performance of CoLB scheme} \label{fig:colb_simulations}
\end{figure}
Fig.~\ref{fig:out_colb} shows the average outage rate of oMSs
when $\delta _f$ is given. As shown in the figure, the average outage
rate of oMSs increases as $\delta _f$ increases.
In our evaluation environments, CoLB cannot expand the service
coverage when the number of fBSs is $30$ or $50$ because the average outage
rate exceeds the configured threshold, i.e., $O_{\max} = 0.15$, even when
$\delta_f$ is $0$~dB. A small margin for coverage expansion is available
when the number of fBSs is $10$. On the other hand, the proposed scheme
adaptively manages the average outage rate according to the
number of fBSs.
Fig.~\ref{fig:th_colb} shows the average throughput
of mMS when the optimal $\delta_f$ is applied to CoLB.
To give more flexibility to coverage control, we consider
a relaxed coverage requirement where $O_{\max} = 0.3$ as well as the basic
requirement, i.e., $O_{\max} = 0.15$.
With $O_{\max} = 0.15$, the performance of CoLB is very close to that of
CoRSSI due to the limited offloading gain,
and only a small performance gain is observed when the number of fBSs is $10$.
On the other hand, a significant performance gain is provided to mMS
with the proposed schemes.
When a relaxed average outage requirement is applied, i.e., $O_{\max} = 0.3$,
CoLB achieves a larger mMS throughput than CoRSSI for most cases
at the cost of increased outage rate. However, the performance gains of mMSs
achieved by the proposed schemes in the relaxed outage requirement
are much more significant than the one achieved by CoLB.
Note that this result does not mean that the proposed scheme
is always better than the co-channel based schemes.
The proposed schemes have strength in improving
the performance of mMSs while limiting the relative
benefits of fMSs and oMSs around the planned levels.
On the other hand, in the aspect of the total system capacity,
the co-channel based scheme is more efficient.
Therefore, the choice of the resource management should
be adaptive to various aspects, e.g.,
the system environments, status of market, consumer's characteristics,
mobile operator's policy, and so on.
\subsection{Impact of Other Environmental Parameters}
\label{sec:other}
Our basic system model used in the previous sections
includes some simplifying assumptions, e,g, uniform user distribution,
to ensure the numerical tractability.
The numerical analysis
and optimization are important in spite of the simplification
because it gives the intuition for the system performance
and optimal parameters.
However, investigation for realistic environments would be beneficial
for further understandings.
In this section, we consider some more environmental
parameters which have not been considered in the basic
numerical model,
and the impacts of the new environmental parameters
are analyzed through the simulations.
\begin{figure*}
\centering \subfigure[Average throughput of mMS with heterogeneous user distribution.]{
\includegraphics[width=0.3\textwidth]{./Hetero_mMS_TH.eps}\label{fig:HETERO_TH_m_M}}\hfill
\subfigure[Limitation from the number of admissible users in an fBS.]{
\includegraphics[width=0.3\textwidth]{./UE_limit_mMS_TH.eps}\label{fig:UE_LIMIT}}\hfill
\subfigure[Simulation results based on real deployment data.]{
\includegraphics[width=0.3\textwidth]{./constructed_seoul.eps}\label{fig:Seoul}}
\caption{Impact of Other Environmental Parameters.} \label{fig:other_parameters}
\end{figure*}
We assume the uniform user distribution in the basic model, but
it is generally said that indoor user density is
much larger than outdoor. Therefore, we introduce new
environmental parameter to represent the
heterogeneity for indoor and outdoor user densities, i.e., $k_{in}$.
We refer to the outdoor and indoor user densities as $\lambda _{u,o}$ and
$\lambda _{u,i}=k_{in}\lambda _{u,o}$, respectively.
Then, the average number of mMSs is obtained by
\begin{equation}
\label{eq:N_m_hetero}
\overline N _m = \left( {A_m - \lambda _f A_m x} \right)\lambda _{u,o},
\end{equation}
and the average number of oMSs in a femtocell area is given by
\begin{equation}
\label{eq:N_o_hetero}
\overline N _o = k_{in} \lambda _{u,o} x\left( {D_h } \right) + \lambda _{u,o} \left( {x - x\left( {D_h } \right)} \right),
\end{equation}
where $x\left( {D_h } \right) = \frac{{1 - e^{ - \pi D_h^2 \lambda _f } }}{{\lambda _f }}$ from (\ref{eq:x}).
By putting (\ref{eq:N_m_hetero}) and (\ref{eq:N_o_hetero}) into
the analysis in Section~\ref{sec:analysis},
the performance of the proposed schemes
with a heterogeneous user distribution can be numerically analyzed.
Fig.~\ref{fig:HETERO_TH_m_M} shows the average throughput of mMS
with a heterogeneous user distribution.
As shown in Fig.~\ref{fig:HETERO_TH_m_M}, the load balancing gain of all the
open access schemes increase as
the indoor user density increases because indoor femtocells
can efficiently offload the traffic of the indoor users.
The relative gain of OA scheme is reduced when the indoor user density is high,
because the comparing schemes also enjoy high offloading gain
from indoor macrocell users. On the other hand, HA scheme still
maintains the relative gain for mMS by limiting
the performance of oMSs.
In the real environments, the maximum number of users served
in a single femtocell is limited due to the capability
of the cheap femtocell device. We refer to
the maximum number of users instantaneously allowed
to access a femtocell as $N_{\max}$.
We investigate the
impact of $N_{\max}$ using simulations.
In the simulations, each macrocell user is dropped in
a random location, and it first tries to choose the best femtocell or
macrocell based on the criterion of the employed target cell selection scheme.
If the target cell is a femtocell and the number of users in the cell
already exceeds $N_{\max}$, then the user tries to choose the next best cell
according to the cell selection criterion.
Multiple simulations are conducted for the quantized target femtocell service area candidates, i.e., $x$. Then, the optimal bandwidth division ratio and intra-femtocell
resource dedicated ratio at a given $x$, i.e., $\rho^*\left(x \right)$ and $\beta^*\left(x \right)$, are numerically obtained by using Proposition~\ref{prop:optimal_bw} and Proposition~\ref{prop:optimal_hybrid}.
Finally, the near optimal $x^*$ is chosen by comparing
the results for various $x$ values.
Fig.~\ref{fig:UE_LIMIT} shows the average throughput of an mMS
when the number of admissible users is limited by a specified value.
If only a small number of users can be served by an fBS,
the average throughput of an mMS cannot be enhanced much although
the load balancing schemes are used.
In our evaluation environments,
the offloading gain sharply decreases when the number of admissible users is less than $6$.
On the other hand, if the number of admitted users is larger than $8$, the performance
enhancement is almost saturated.
\label{sec:building}
Our basic assumption for femtocell distribution
is that the buildings~(or houses) are
randomly distributed according to SPPP.
In order to show the performance in
the realistic environments,
we show the simulation results based on the real
deployment information for WiFi APs.
Because the femtocell service is in an early phase,
we use the location information of WiFi APs registered
in~\cite{wigle}, where we assume
that some of the WiFi APs are replaced by femtocells.
Simulation results shown in Fig.~\ref{fig:Seoul} have been obtained
based on the deployment information around the Seoul station, which
is the downtown of Seoul, Korea. In order to compare the results with
the basic results in Section~\ref{sec:simulations}, we assume that only the predefined ratio
of WiFi APs in the area are replaced by femtocells, and the density of the femtocell in
the macrocell area is known to the system. We obtain the averaged value after
$300$ simulations where different randomly chosen WiFi APs' locations
are assumed in each simulation run. The same channel model
which is used in the numerical analysis is used for the simulations.
The meaning of this simulation result may be limited because the performance
gap between the numerical analysis and
real performance results is severely affected by the system environment.
However, the simulation results in Fig.~\ref{fig:Seoul}
shows that analysis results show similar trends to the simulation in some cases,
and the proposed schemes still have performance gain even though the
optimization based on the local information has not been performed.
Although the analysis results can be different from the performance in the real system,
the numerical analysis and optimization are still very important because it can give
the insights for the network performance in the various environments and good guidelines for the
system design.
\section{Conclusion}
\label{sec:conclusion}
In this paper, we develop the load balancing schemes which are efficient in
the environments where the open or hybrid access femtocells coexist with the macrocells.
We aim at maximizing the average throughput
performance of
mMSs while guaranteeing some amount of benefits to the
fMSs who deploy femtocells in their homes.
In order to maximize the offloading gain of the open and hybrid access femtocells, we propose
to use the separate bandwidth deployment, and we jointly optimize the
effective service areas of femtocells
and the amount of dedicated resources for femtocells.
Using the analytic model, we prove that the joint optimization problem is a convex optimization problem
in some typical environments.
We also introduce the scheme which applies the interference thinning scheme on top of the proposed schemes
in order to further enhance the offloading gain.
In the hybrid access femtocells, the performance of our load balancing scheme is
improved by optimally determining the amount of dedicated resources only for the fMSs.
Performance evaluation results show that the proposed schemes significantly improve the system-wide
performance while satisfying the requirements of fMSs.
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Recent results from the LHC and direct dark matter detection experiments
constrain considerably possible scenarios beyond the Standard Model
(SM), amongst others its supersymmetric (SUSY) extensions. These
constraints originate essentially from the Higgs mass~\cite{Aad:2012tfa,
Chatrchyan:2012ufa} and its quite SM like signal rates, the absence of
signals in searches for squarks and gluinos after the 8~TeV run at the
LHC~\cite{TheATLAScollaboration:2013fha, Chatrchyan:2014lfa}, and
upper bounds on dark matter--nucleus cross sections from the LUX
experiment~\cite{Akerib:2013tjd}.
Masses and couplings of Higgs boson(s), SUSY particles and notably the
lightest SUSY particle (LSP, the dark matter candidate), are strongly
correlated in SUSY extensions of the SM if one assumes at least
partial unification of the soft SUSY breaking terms at a grand
unification (GUT) scale.
Hence it is interesting to study how the combined constraints affect the
parameter space and, notably, which signals beyond the SM we can expect
in the future. Such studies (after the discovery of the 126~GeV Higgs
boson) had been performed earlier in the Minimal SUSY extension of the
SM (MSSM)~\cite{ Baer:2011ab, Arbey:2011ab, Buchmueller:2011ab,
Akula:2011aa, Kadastik:2011aa, Cao:2011sn, Ellis:2012aa, Fowlie:2012im,
Beskidt:2012sk, Buchmueller:2012hv, Strege:2012bt, Ellis:2012nv,
Cabrera:2012vu, Kowalska:2013hha, Cohen:2013kna, Beskidt:2013gia,
Henrot-Versille:2013yma, Bechtle:2013mda, Kim:2013uxa,
Buchmueller:2013rsa, Ellis:2013oxa} and the Next-to-Minimal SUSY
extension of the SM (NMSSM)~\cite{Arbey:2011ab, Gunion:2012zd,
Ellwanger:2012ke, Belanger:2012tt, Kowalska:2012gs,
Kim:2013uxa,Beskidt:2014oea}.
These studies differ, however, in the treatment of the soft SUSY
breaking terms in the Higgs sector at the GUT scale: In ``fully
constrained'' versions of the MSSM or NMSSM these are supposed to be
unified with the soft SUSY breaking terms in the squark and slepton
sectors. In NUHM (non-universal Higgs masses) or ``semi-constrained''
versions of the MSSM or NMSSM one allows the soft SUSY breaking terms in
the Higgs sector to be different; after all the quantum numbers of the
Higgs fields differ from those of quarks and leptons: Higgs fields are
in a real representation ($2 + \bar{2}$) of SU(2), but do not fit into
complete representations of SU(5); these properties can easily have an
impact on the presently unknown sources of soft SUSY breaking terms. In
the NMSSM including the singlet superfield $S$, ``semi-constrained'' can
indicate non-universal soft SUSY breaking terms involving the singlet
only, or non-universal soft SUSY breaking terms involving SU(2) doublet
or singlet Higgs fields. In the present study we allow for the latter
more general case.
Previous studies of the NMSSM with constraints at the GUT scale~\cite{
Gunion:2012zd, Ellwanger:2012ke, Belanger:2012tt, Kowalska:2012gs,
Kim:2013uxa, Beskidt:2014oea} had found that wide ranges of parameter
space comply with constraints from the LHC and on dark matter, and that
less ``tuning'' is required than in the MSSM~\cite{Kowalska:2012gs,
Kim:2013uxa}. These findings are confirmed by scans of the parameter
space of the general NMSSM (without constraints at the GUT scale)~\cite{
Kang:2012sy, Cao:2012fz, Vasquez:2012hn, Perelstein:2012qg,Agashe:2012zq,
Gherghetta:2012gb, Cheng:2013fma}, and motivate a thorough analysis of
the semi-constrained NUH-NMSSM with up-to-date experimental constraints,
amongst others on Higgs signal rates and bounds on dark matter--nucleus
cross sections~\cite{Akerib:2013tjd}. ``NUH'' appears without ``M''
since, apart from the Higgs mass terms, also trilinear couplings
involving Higgs bosons only are allowed to differ from trilinear
couplings involving squarks or sleptons at the GUT scale, see the next
section.
Using the code {\sf NMSPEC}~\cite{Ellwanger:2006rn} within {\sf
NMSSMTools\_4.2.1}~\cite{Ellwanger:2004xm,Ellwanger:2005dv}
together with {\sf micr}\-{\sf OMEGAS\_3} \cite{Belanger:2013oya} we
have sampled about 3.2~M viable points in the
parameter space, which allows us to cover the complete range of masses
and couplings of the LSP and additional Higgs bosons, parts of which had
not been observed in previous analyses. In this paper we confine
ourselves to regions where an additional NMSSM-specific Higgs scalar is
lighter than the SM-like Higgs boson near 126~GeV; this region is
strongly favoured by the mass of the SM-like Higgs boson, and contains
the most interesting phenomena to be searched for in the future.
In the next section we present the model, the applied phenomenological
constraints, the definition of fine-tuning, and the ranges of
parameters scanned over. In section~3 we discuss the impact of
unsuccessful searches for squarks and gluinos at the LHC on fine-tuning
and some of the parameters like the soft squark/slepton masses $m_0$,
the universal gaugino masses $M_{1/2}$ and the NMSSM-specific Yukawa
coupling $\lambda$. Section~4 is devoted to the properties of the LSP,
its detection rates to be expected in the future, and its annihilation
processes allowing for a viable relic density. In section~5 we discuss
the Higgs sector, in particular prospects to detect the lighter NMSSM
specific Higgs scalar. Conclusions and an outlook are given in
section~6.
\section{The NMSSM with constraints at the GUT scale}
The NMSSM~\cite{Ellwanger:2009dp} differs from the MSSM due to the
presence of the gauge singlet superfield $S$. In the simplest $\mathbb
Z_3$ invariant realisation of the NMSSM, the Higgs mass term $\mu H_u
H_d$ in the superpotential $W_\mathrm{MSSM}$ of the MSSM is replaced by
the coupling $\lambda$ of $S$ to $H_u$ and $H_d$ and a self-coupling
$\kappa S^3$. Hence, in this simplest version the superpotential
$W_\mathrm{NMSSM}$ is scale invariant and given by
\begin{equation}} \def\eeq{\end{equation}\label{eq:2.1}
W_\mathrm{NMSSM} = \lambda \hat S \hat H_u\cdot \hat H_d +
\frac{\kappa}{3}
\hat S^3 + \dots\; ,
\eeq
where hatted letters denote superfields, and the ellipses denote the
MSSM-like Yukawa couplings of $\hat H_u$ and $\hat H_d$ to the
quark and lepton superfields. Once the real scalar component of
$\hat S$ develops a vev $s$, the first term in $W_\mathrm{NMSSM}$
generates an effective $\mu$-term
\begin{equation}} \def\eeq{\end{equation}\label{eq:2.2}
\mu_\mathrm{eff}=\lambda\, s\; .
\eeq
The soft Susy breaking terms consist of mass terms for the Higgs bosons
$H_u$, $H_d$, $S$, squarks
$\tilde{q_i} \equiv (\tilde{u_i}_L, \tilde{d_i}_L$), $\tilde{u_i}_R^c$,
$\tilde{d_i}_R^c$ and sleptons $\tilde{\ell_i} \equiv (\tilde{\nu_i}_L,
\tilde{e_i}_L$) and $\tilde{e_i}_R^c$
(where $i=1,2,3$ is a generation index):
\begin{eqnarray}} \def\eea{\end{eqnarray}
-{\cal L}_\mathrm{0} &=&
m_{H_u}^2 | H_u |^2 + m_{H_d}^2 | H_d |^2 +
m_{S}^2 | S |^2 +m_{\tilde{q_i}}^2|\tilde{q_i}|^2
+ m_{\tilde{u_i}}^2|\tilde{u_i}_R^c|^2
+m_{\tilde{d_i}}^2|\tilde{d_i}_R^c|^2\nonumber \\
&& +m_{\tilde{\ell_i}}^2|\tilde{\ell_i}|^2
+m_{\tilde{e_i}}^2|\tilde{e_i}_R^c|^2\; ,
\label{eq:2.3}
\eea
trilinear interactions involving the third generation squarks, sleptons and the
Higgs fields (neglecting the Yukawa couplings of the two first generations):
\begin{eqnarray}} \def\eea{\end{eqnarray}
-{\cal L}_\mathrm{3}&=&
\Bigl( h_t A_t\, Q\cdot H_u \: \tilde{u_3}_R^c +
h_b A_b\, H_d \cdot Q \: \tilde{d_3}_R^c +
h_\tau A_\tau \,H_d\cdot L \: \tilde{e_3}_R^c \nonumber \\
&& +\, \lambda A_\lambda\, H_u \cdot H_d \,S + \frac{1}{3} \kappa
A_\kappa\, S^3 \Bigl)+ \, \mathrm{h.c.}\;,
\label{eq:2.4}
\eea
and mass terms for the gauginos $\tilde{B}$ (bino), $\tilde{W}^a$
(winos) and $\tilde{G}^a$ (gluinos):
\begin{equation}} \def\eeq{\end{equation}\label{eq:2.5}
-{\cal L}_\mathrm{1/2}= \frac{1}{2} \bigg[
M_1 \tilde{B} \tilde{B}
\!+\!M_2 \sum_{a=1}^3 \tilde{W}^a \tilde{W}_a
\!+\!M_3 \sum_{a=1}^8 \tilde{G}^a \tilde{G}_a
\bigg]+ \mathrm{h.c.}\;.
\eeq
In constrained versions of the NMSSM one assumes that the soft Susy
breaking terms involving gauginos, squarks or sleptons are identical at
the GUT scale:
\begin{equation}} \def\eeq{\end{equation} \label{eq:2.6}
M_1 = M_2 = M_3 \equiv M_{1/2}\; ,
\eeq
\begin{equation}} \def\eeq{\end{equation} \label{eq:2.7}
m_{\tilde{q_i}}^2= m_{\tilde{u_i}}^2=m_{\tilde{d_i}}^2=
m_{\tilde{\ell_i}}^2=m_{\tilde{e_i}}^2\equiv m_0^2\; ,
\eeq
\begin{equation}} \def\eeq{\end{equation} \label{eq:2.8}
A_t = A_b = A_\tau \equiv A_0\; .
\eeq
In the NUH-NMSSM considered here one allows the Higgs sector to play a
special role: The Higgs soft mass terms $m_{H_u}^2$, $m_{H_d}^2$ and
$m_{S}^2$ are allowed to differ from $m_0^2$ (and determined implicitely
at the weak scale by the three minimization equations of the effective potential),
and the trilinear
couplings $A_{\lambda}$, $A_{\kappa}$ can differ from $A_0$. Hence the
complete parameter space is characterized by
\begin{equation}} \def\eeq{\end{equation} \label{eq:2.9}
\lambda\ , \ \kappa\ , \ \tan\beta\ ,\ \mu_\mathrm{eff}\ , \ A_{\lambda}
\ , \ A_{\kappa} \ , \ A_0 \ , \ M_{1/2}\ , \ m_0\; ,
\eeq
where the latter five parameters are taken at the GUT scale.
Expressions for the mass matrices of the physical CP-even and CP-odd
Higgs states -- after $H_u$, $H_d$ and $S$ have assumed vevs $v_u$,
$v_d$ and $s$ and including the dominant radiative corrections -- can be
found in~\cite{Ellwanger:2009dp} and will not be repeated here. The
physical CP-even Higgs states will be denoted as $H_i$, $i=1,2,3$
(ordered in mass), and the physical CP-odd Higgs states as $A_i$,
$i=1,2$. The neutralinos are denoted as $\chi^0_i$, $i=1,...,5$ and
their mixing angles $N_{i,j}$ such that $N_{1,5}$ indicates the singlino
component of the lightest neutralino $\chi^0_1$.
Subsequently we are interested in regions of the parameter space where
doublet-singlet mixing in the Higgs sector leads to an increase of the
mass of the SM-like (mostly doublet-like) Higgs boson, which leads
naturally to a SM-like Higgs boson $H_2$ in the 126~GeV range~\cite{
Hall:2011aa,Ellwanger:2011aa,Arvanitaki:2011ck,King:2012is, Kang:2012sy,
Cao:2012fz}, but implies a lighter mostly singlet-like Higgs state
$H_1$.
Recent phenomenological constraints include, amongst others, upper
bounds on the direct (spin independent) detection rate of dark matter by
LUX~\cite{Akerib:2013tjd}. In the NMSSM, the LSP (the dark matter candidate)
is assumed to be the lightest neutralino, as in the MSSM.
Its spin independent detection rate and relic
density are computed with the help of
{\sf micrOMEGAS\_3} ~\cite{Belanger:2013oya}.
We apply the upper bounds of LUX and
require a relic density inside a slightly enlarged WMAP/Planck
window~\cite{Hinshaw:2012aka,Ade:2013zuv} $0.107 \leq \Omega h^2 \leq
0.131$ in order not to loose too many points in parameter space; the
precise value of $\Omega h^2$ has little impact on the subsequent
results.
In the Higgs sector we require a neutral CP-even state with a mass of
$125.7 \pm 3$~GeV allowing for theoretical and parametric uncertainties
of the mass calculation; we used 173.1~GeV for the top quark mass. Its
signal rates should comply with the essentially SM-like signal rates in
the channels measured by ATLAS/CMS/Teva\-tron. These measurements can be
combined leading to 95\% confidence level (CL) contours in the planes of
Higgs production via (gluon fusion and ttH) -- (vector boson fusion and
associate
production with W/Z), separately for Higgs decays into $\gamma\gamma$,
ZZ or WW and $b\bar{b}$ or $\tau^+\tau^-$. We require that the signal
rates for a Higgs boson in the above mass range are within all three
95\% confidence level contours derived in~\cite{Belanger:2013xza}.
The application of constraints from unsuccessful searches for sparticles
at the first run of the LHC is more delicate: These bounds depend on all
parameters of the model via the masses and couplings (and the resulting
decay cascades) of all sparticles. However, it is possible to proceed as
follows, using the most constraining searches for gluinos and squarks of
the first generation in events with jets and missing $E_T$: For heavy
squarks and/or gluinos the production cross sections are so small that
these points in parameter space are not excluded independently of the
squark/gluino decay cascades. On the other hand, relatively light
squarks and/or gluinos are excluded independently of their decay
cascades. In between these regions defined in the planes of
squark/gluino masses or $m_0/M_{1/2}$, exclusion does depend on their
decays, in particular on the presence of a light singlino-like LSP at
the end of the cascades~\cite{Das:2012rr,Das:2013ta}.
The boundaries between these three regions were obtained with
the help of the analysis of some hundreds of points in parameter space:
Events were generated by MadGraph/MadEvent \cite{Alwall:2011uj} which
includes Pythia~6.4~\cite{Sjostrand:2006za} for showering and
hadronisation. The sparticle branching ratios are obtained with the help
of the code NMSDECAY~\cite{Das:2011dg} (based on
SDECAY~\cite{Muhlleitner:2003vg}), and are passed to Pythia. The output in
StdHEP format is given to CheckMATE~\cite{Drees:2013wra} which includes
the detector simulation DELPHES~\cite{deFavereau:2013fsa} and compares
the signal rates to constraints in various search channels of ATLAS and
CMS. Corresponding results will be presented in section~3.
Other constraints from $b$-physics, LEP (from Higgs searches and
invisible Z decays) and the LHC (on heavy Higgs bosons decaying into
$\tau^+ \tau^-$) are applied as in
{\sf NMSSMTools\_4.2.1}~\cite{Ellwanger:2004xm,Ellwanger:2005dv},
leaving aside the muon anomalous magnetic moment.
Since the fundamental parameters of the model are the masses and
couplings at the GUT scale, it makes sense to ask in how far these have
to be tuned relative to each other in order to comply with the SM-like
Higgs mass and the non-observation of sparticles at the LHC. To this end
we consider the usual measure of fine-tuning~\cite{Barbieri:1987fn}
\begin{equation}} \def\eeq{\end{equation}\label{eq:2.10}
FT = Max\left\{\left|\frac{\partial \ln(M_Z)}{\partial \ln(p_i^{GUT})}
\right|\right\}
\eeq
where $p_i^{GUT}$ denote all dimensionful and dimensionless parameters
(Yukawa couplings, mass terms and trilinear couplings) at the GUT scale.
$FT$ is computed numerically in {\sf NMSSMTools\_4.2.1} following the
method described in~\cite{Ellwanger:2011mu} where details can be found.
We have scanned the parameter space of the NUH-NMSSM given
in~(\ref{eq:2.9}) using a Markov Chain Monte Carlo (MCMC) technique. In
addition to the phenomenological constraints discussed above we require
the absence of Landau singularities of the running Yukawa couplings
below the GUT scale, and the absence of deeper unphysical minima of the
Higgs potential with at least one vanishing vev $v_u$, $v_d$ or $s$.
Bounds on the dimensionful parameters follow from
the absence of too large fine-tuning; we imposed $FT < 1000$. Finally we
obtained $\sim 3.2\times10^6$ valid points in parameter space within the
following ranges of the parameters~(\ref{eq:2.9}):
\be
1\times10^{-6} \leq \lambda \leq 0.722,& -0.08 \leq \kappa \leq 0.475,&
1.42 \leq \tan\beta \leq 60.3,\nonumber\\
-537\ \mathrm{GeV} \leq \mu_\mathrm{eff} \leq 753\ \mathrm{GeV},&
-19\ \mathrm{TeV} \leq A_{\lambda} \leq 8.5\ \mathrm{TeV},
&-1.3\ \mathrm{TeV} \leq A_{\kappa} \leq 5.3\ \mathrm{TeV},\nonumber\\
0\ \leq m_0 \leq 4.4\ \mathrm{TeV},& 0.1\ \mathrm{TeV}
\leq M_{1/2} \leq 3.1\ \mathrm{TeV},&
-6.6\ \mathrm{TeV} \leq A_{0} \leq 8.1\ \mathrm{TeV}.\nonumber\\
&&\label{eq:2.11}
\eea
The fact that the upper bounds on the dimensionful parameters are
distinct originates from the different impact of these parameters on the
fine-tuning, which is often dominated by the universal gaugino mass
parameter $M_{1/2}$.
\section{Impact of LHC constraints on squark/gluino masses and
fine-tuning}
Strong constraints on parameter spaces of SUSY extensions of the SM come
from searches for gluinos $\tilde{g}$ and squarks $\tilde{q}$ of the
first generation in events with jets and missing
$E_T$~\cite{TheATLAScollaboration:2013fha, Chatrchyan:2014lfa}.
In~\cite{TheATLAScollaboration:2013fha} exclusion limits for MSUGRA/CMSSM
models have been given in the $m_0-M_{1/2}$ and
$M_{\tilde{g}}-m_{\tilde{q}}$ planes for $\tan\beta=30$, $A_0=-2m_0$ and
$\mu>0$.
As a result of the simulations described in the previous section we
found that the 95\%~CL upper limits on signal events
in~\cite{TheATLAScollaboration:2013fha} lead to exclusion limits in the
$m_0-M_{1/2}$ or $M_{\tilde{g}}-m_{\tilde{q}}$ planes in the NUH-NMSSM
which are very similar to the CMSSM if the LSP is bino-like, but can be
alleviated in the presence of a light singlino-like LSP at the end of
the cascades~\cite{Das:2012rr,Das:2013ta}. Still, even with a
singlino-like LSP, certain regions in these planes are always excluded.
\begin{figure}[ht!]\centering
\includegraphics[width=.5\textwidth]{fig1a.png}
\hspace*{-2em}
\includegraphics[width=.5\textwidth]{fig1b.png}
\vspace*{-2em}
\caption{The $m_0-M_{1/2}$ and
$M_{\tilde{g}}-m_{\tilde{q}}$ planes in the NUH-NMSSM.
Green: regions allowed by the 95\% CL upper limits on signal events
in~\cite{TheATLAScollaboration:2013fha}, blue: regions allowed in
the presence of a singlino-like LSP, red: regions which are
always excluded.}
\label{fig:1}
\end{figure}
In Fig. \ref{fig:1} we show the $m_0-M_{1/2}$ and
$M_{\tilde{g}}-m_{\tilde{q}}$ planes in the NUH-NMSSM and indicate in
green the regions allowed by the 95\% CL upper limits on signal events
(practically identical to the ones given
in~\cite{TheATLAScollaboration:2013fha}), in blue the regions possibly
allowed in the presence of a singlino-like LSP, and in red the regions
which are always excluded. Note that, in contrast to the MSSM, the limit
$m_0 \to 0$ is always possible for all $M_{1/2}$: In the MSSM this
region is limited by the appearance of a stau LSP. In the NMSSM a
singlino-like LSP can always be lighter than the lightest stau, and its
relic density can be reduced to the WMAP/Planck window through
singlino-stau coannihilation as in the fully constrained
NMSSM~\cite{Djouadi:2008yj,Djouadi:2008uj} or through narrow resonances
implying specific NMSSM light Higgs states~\cite{Belanger:2005kh,
Hugonie:2007vd, Belanger:2008nt, Vasquez:2012hn}. (The combined
constraints from the Higgs sector and the nature of the LSP lead to
discontinuities in the allowed parameter space for small $m_0$.)
These lower bounds on the squark and gluino masses dominate the lower
bounds on the fine-tuning $FT$ defined in (\ref{eq:2.10}). In
Fig.~\ref{fig:2} we show $FT$ as function of the squark and gluino
masses, and the impact of the LHC constraints in the same color coding
as in Fig.~\ref{fig:1}.
\begin{figure}[ht!]\centering
\includegraphics[width=.5\textwidth]{fig2a.png}
\hspace*{-2em}
\includegraphics[width=.5\textwidth]{fig2b.png}
\vspace*{-2em}
\caption{$FT$ as defined in (\ref{eq:2.10}) as function of the squark
and gluino masses, and the impact of the LHC constraints in the same
color coding as in Fig. \ref{fig:1}.}
\label{fig:2}
\end{figure}
We see that the LHC forbidden red region increases the lower bound on
$FT$ from $\sim 20$ to $FT \;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 80$; the NMSSM-specific alleviation
(blue region) has a minor impact on $FT$. The dominant contribution to
$FT$ in (\ref{eq:2.10}) originates typically from $M_{1/2}$ (i.e. the
gluino mass at the GUT scale), or from the soft Higgs mass term
$m_{H_u}^2$. If one requires unification of $m_{H_u}$ and $m_{H_d}$ with
$m_0$ as in~\cite{Kowalska:2012gs}, $FT$ is considerably larger $(\;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\;
400)$. In the MSSM -- after imposing LHC constraints on squark and
gluino masses, defining $FT$ with respect to parameters at the GUT scale
and allowing for non-universal Higgs mass terms at the GUT scale as
in~\cite{Baer:2012cf} -- one finds $FT \;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 1000$. The much lower value of
$FT$ in the NUH-NMSSM coincides with the result in~\cite{Binjonaid:2014oga}.
The impact of $M_{1/2}$ on $FT$ is actually indirect: Heavy gluinos lead
to large radiative corrections to the stop masses which,
in turn, lead to large radiative corrections to the soft Higgs mass
terms. Therefore, if one defines $FT$ with respect to parameters at a
lower scale, low $FT$ is typically related to light stops. On the
left-hand side of Fig.~\ref{fig:3} we show $FT$ as function of the mass
$m_{\tilde{t}_1}$ of the lightest stop. We see that, without imposing LHC
constraints on squark and gluino masses, the lower bound on $FT$ (still
with respect to parameters at the GUT scale) would increase slightly with
$m_{\tilde{t}_1}$, but with LHC constraints the lower bound on $FT$
depends weakly on (decreases only slightly with) $m_{\tilde{t}_1}$.
\begin{figure}[ht!]\centering
\includegraphics[width=.5\textwidth]{fig3a.png}
\hspace*{-2em}
\includegraphics[width=.5\textwidth]{fig3b.png}
\vspace*{-2em}
\caption{Left: $FT$ as function of the $m_{\tilde{t}_1}$. Right: $FT$ as
function of $\lambda$ at the SUSY scale. The
color coding is as in Fig.~\ref{fig:1}.}
\label{fig:3}
\end{figure}
In the MSSM, the measured mass of the SM-like Higgs $H_{SM}$ requires relatively
heavy stops and/or a Higgs-stop trilinear coupling $A_t$, which also
contribute to $FT$. In the NMSSM (recall that, in the scenario considered
here, $H_{SM}=H_2$) large radiative corrections to the
SM-like Higgs mass $m_{H_{SM}}$ are not required, since the SM-like
Higgs mass can be pushed upwards either through a positive tree level
contribution $\sim \lambda^2 \sin^2 2\beta$~\cite{Ellwanger:2009dp}, or
through mixing with a lighter Higgs state $H_1$~\cite{Ellwanger:1999ji}
which does not require large values of $\lambda$~\cite{Badziak:2013bda}.
(In the latter scenario too
large values of $\lambda$, i.e. a too large $H_1 - H_{SM}$ mixing angle,
can imply an inacceptable reduction of the signal rates of $H_{SM}$ at
the LHC and/or lead to the violation of LEP constraints on $H_1$.)
On the right-hand side of Fig.~\ref{fig:3} we show $FT$ as function of
$\lambda$. We see that -- without imposing LHC constraints on squark and
gluino masses -- the minimum of $FT$ would indeed be assumed for
$\lambda \sim 0.6$ related to the tree level contribution $\sim
\lambda^2 \sin^2 2\beta$ to $m_{H_{SM}}$. Including LHC constraints,
local minima of $FT$ exist both for $\lambda \approx 0.6$ and $\lambda
\approx 0.1$.
Since the increase of the SM-like Higgs mass with the help of the tree
level contribution $\sim \lambda^2 \sin^2 2\beta$ is effective only for
large $\lambda$ but relatively low $\tan\beta$, these regions are
typically correlated which is clarified on the left hand side of
Fig.~\ref{fig:4}. On the right hand side we show the correlations
between $\lambda$ and $\kappa$ which shows that larger $\kappa$ are
typically related to larger~$\lambda$.
\begin{figure}[ht!]\centering
\includegraphics[width=.5\textwidth]{fig4a.png}
\hspace*{-2em}
\includegraphics[width=.5\textwidth]{fig4b.png}
\vspace*{-2em}
\caption{Left: $\lambda$ as function of $\tan\beta$. Right: $\kappa$ as
function of $\lambda$. The colors are as in Fig.~\ref{fig:1}.}
\label{fig:4}
\end{figure}
Herewith we conclude the discussion of the impact of LHC constraints on
$FT$ and the corresponding correlations with other parameters.
\section{Properties of dark matter}
Besides the enlarged Higgs sector, the enlarged neutralino sector of the
NMSSM can have a significant phenomenological impact. The LSP
(the lightest neutralino $\chi_1^0$) can have a dominant singlino
component and still be an acceptable candidate for dark matter. Its
relic density can be reduced to fit in the WMAP/Planck window, amongst
others, via the exchange of NMSSM-specific CP-even or CP-odd Higgs
scalars in the s-channel~\cite{Belanger:2005kh, Hugonie:2007vd,
Belanger:2008nt, Vasquez:2012hn}, whereas its direct detection cross
section can be very small.
The latter feature is clarified in Fig.~\ref{fig:5} where we show
the spin-independent $\chi_1^0$-nucleon cross section (after imposing
constraints from the LUX experiment~\cite{Akerib:2013tjd}) as function
of $M_{\chi_1^0}$. We
focus on $\chi_1^0$ masses below 100~GeV since no additional interesting
features appear for larger $M_{\chi_1^0}$, but the region of small
$M_{\chi_1^0}$ exhibits structures which ask for explanations.
In Fig.~\ref{fig:5} we have indicated the expected neutrino background
to future direct dark matter detection experiments from~\cite{Billard:2013qya}
as a black line; it will be difficult to
impossible to measure $\chi_1^0$-nucleon cross section smaller than this
background. Unfortunately we see that significant regions in the
NUH-NMSSM parameter space -- notably for $M_{\chi_1^0} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 10$~GeV or
$M_{\chi_1^0} \;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 60$~GeV -- may lead to such small cross sections.
\begin{figure}[ht!]\centering
\includegraphics[width=.7\textwidth]{fig5.png}
\vspace*{-2em}
\caption{The spin-independent $\chi_1^0$-nucleon cross section
$\sigma_{\mathrm SI}$ (after
imposing constraints from the LUX experiment~\cite{Akerib:2013tjd}) as
function of $M_{\chi_1^0}$, focussing on $M_{\chi_1^0}<100$~GeV. The
black line indicates the expected neutrino background to future direct
dark matter detection experiments (from~\cite{Billard:2013qya}).
The colors are as in Fig.~\ref{fig:1}.}
\label{fig:5}
\end{figure}
Small $\chi_1^0$-nucleon cross sections originate from a large singlino
component of $\chi_1^0$. Its singlino component $N_{15}^2$ is shown as
function of $M_{\chi_1^0}$ in Fig.~\ref{fig:6}.
\begin{figure}[ht!]\centering
\includegraphics[width=.7\textwidth]{fig6.png}
\vspace*{-2em}
\caption{The $\chi_1^0$ singlino component (squared) as function of
$M_{\chi_1^0}$. The colors are as in Fig.~\ref{fig:1}.}
\label{fig:6}
\end{figure}
Different regions of $M_{\chi_1^0}$ correspond to different dominant
diagrams contributing to $\chi_1^0-\chi_1^0$ annihilation before its
freeze-out. For small $M_{\chi_1^0} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 30$~GeV these are the exchange
of NMSSM-specific CP-even or CP-odd Higgs scalars with masses $\approx 2
M_{\chi_1^0}$ in the s-channel, with couplings originating from the
cubic $S^3$ term proportional to $\kappa$ in the superpotential. For
$M_{\chi_1^0} \sim 40-48$~GeV, $\chi_1^0-\chi_1^0$ annihilation is
dominated by $Z$-exchange in the s-channel. The larger is the singlino
component of $\chi_1^0$, the closer $M_{\chi_1^0}$ has to be to $M_Z/2$
in order to compensate for the smaller coupling. For $M_{\chi_1^0} \sim
55-62$~GeV, $\chi_1^0-\chi_1^0$ annihilation is dominated by
$H_{SM}$-exchange. In the empty regions for $M_{\chi_1^0} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 55$~GeV,
the non-singlet components of $\chi_1^0$ would have to be so large for
successful $\chi_1^0-\chi_1^0$ annihilation that the $\chi_1^0$-nucleon
cross section would violate constraints from LUX. For $M_{\chi_1^0}
\;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 62$~GeV $\chi_1^0$ can have sizeable bino and/or higgsino
components allowing for numerous additional (e.g. MSSM-like)
$\chi_1^0-\chi_1^0$ annihilation channels.
\section{Properties of the lighter Higgs boson $H_1$}
In this paper we focus on scenarios where mixing of the SM-like Higgs
boson $H_{SM}$ with a lighter NMSSM-specific mostly singlet-like Higgs
boson $H_1$ helps to increase the mass of $H_{SM}$. This is possible
even for relatively small values of $\lambda \approx 0.1$ and moderate
to large values of $\tan\beta$~\cite{Badziak:2013bda}.
However, the $H_{SM}-H_1$ mixing angle must not be too large: It leads
to a reduction of the $H_{SM}$ couplings to electroweak gauge bosons and
quarks, hence to a reduction of its production cross section at the LHC.
These must comply with the measured signal rates, for which we require
values inside the 95\% CL contours of~\cite{Belanger:2013xza}. Moreover,
for $M_{H_1} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 114$~GeV, $H_1$ must satisfy constraints from Higgs
searches at LEP~\cite{Schael:2006cr}.
Hence the question is whether there are realistic prospects for the
discovery of $H_1$ at the LHC~\cite{Cacciapaglia:2013ora}. First we
consider the case where $H_1$ does not decay dominatly into pairs of
lighter NMSSM-specific CP-odd Higgs bosons. The branching fractions of
$H_1$ into $ZZ$ and $W^+ W^-$ are small, both due to its smaller mass
and its reduced couplings to $ZZ$ and $W^+ W^-$.
The branching fraction of $H_1$ into $\gamma\gamma$ can be
considerably larger than the one of a SM-like Higgs boson of the same
mass~\cite{Ellwanger:2010nf,Badziak:2013bda}, both due to a possible
reduction of its width into the dominant $b\bar{b}$ channel through
mixing, and/or due to additional (higgsino-like) chargino loops
contributing to the $H_1-\gamma\gamma$ coupling where the latter involve
the NMSSM-specific coupling $\lambda$~\cite{SchmidtHoberg:2012yy,
Choi:2012he}.
However, due to the reduced coupling of $H_1$ to SM particles, its
production cross section $\sigma_{H_1}$ is smaller than the one of a
SM Higgs boson $H^{SM}$ of the same mass. Hence one has to consider
the reduced signal rate $\sigma_{H_1}\times BR(H_1\to\gamma\gamma)/
\left(\sigma_{H^{SM}}\times BR(H^{SM}\to\gamma\gamma
)\right)$~\cite{Ellwanger:2010nf, Cao:2011pg,Badziak:2013bda,King:2012tr}
which is shown for production via gluon fusion in Fig.~\ref{fig:7}.
\begin{figure}[ht!]\centering
\includegraphics[width=.7\textwidth]{fig7.png}
\vspace*{-2em}
\caption{The $H_1$ signal rate in gluon fusion and the $\gamma\gamma$
channel relative to a SM-like Higgs boson $H^{SM}$ of the same
mass. The color code is as in Fig.~\ref{fig:1}.}
\label{fig:7}
\end{figure}
We see that the signal rate can be about 3.5 times larger than the one
of a SM-like Higgs boson of a mass of $\sim 60$~GeV. The absence of
points with large signal rates for $M_{H_1} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 60$~GeV follows from
the constraints on the signal rates of $H_{SM}$: For $M_{H_1} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\;
60$~GeV, $H_{SM}$ could decay into a pair of $H_1$ bosons, and this
decay channel is easily dominant if kinematically allowed. The
corresponding reductions of the other $H_{SM}$ branching fractions would
be incompatible with its measured signal rates. (The possible
enhancement of the signal rate for $M_{H_1}\;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 3.5$~GeV originates
from the absence of decays into $b\bar{b}$ and $\tau^+\tau^-$, which
makes it very sensitive to relative enhancements of the width into
$\gamma\gamma$ via chargino loops.) For $M_{H_1} \;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 110$~GeV, some
points with a reduced signal rate $\;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 0.5$ could actually already be
excluded by limits from CMS in~\cite{CMSgamgam} depending, however, on
the relative contribution of gluon fusion to the expected signal rate in
this mass range. On the other hand it is clear that, for $M_{H_1} \sim
M_Z$, the $H_1\to\gamma\gamma$ channel faces potentially large
backgrounds from fake photons from $Z\to e^+ e^-$~decays.
For the $b\bar{b}$ and $\tau^+\tau^-$ final states we found that due to
the reduction of the production cross section and the reduction of the
couplings (i.e. branching fractions) of $H_1$ its reduced signal rates
in gluon fusion, vector boson fusion and associate production with $Z/W$
are always below 0.3 for $M_{H_1} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 114$~GeV, and still below 0.6
for $114\ \mathrm{GeV}\ \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; M_{H_1} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 126$~GeV; hence we will not
further analyse these channels (also plagued by the absence of narrow
peaks in the invariant mass of the final states).
Another possibility is that $H_1$ decays dominantly into pairs of light
NMSSM-specific CP-odd Higgs bosons $A_1$ (see~\cite{Cerdeno:2013cz} and
refs. therein). If this channel is open, the corresponding branching
fraction $BR(H_1 \to A_1 A_1)$ can vary from 0 to 1 for all $M_{H_1}$
and $M_{A_1}$. However, the production cross section of $H_1$ is always
reduced relative to the one of a SM-like Higgs boson $H^{SM}$ of the
same mass. Focussing again on gluon fusion, we show in Figs.~\ref{fig:8}
the $BR(H_1 \to A_1 A_1)$ multiplied by the reduced $H_1$ production
cross section (relative to the one of a SM-like Higgs boson $H^{SM}$ of
the same mass) as function of $M_{H_1}$ and $M_{A_1}$.
\begin{figure}[ht!]\centering
\includegraphics[width=.5\textwidth]{fig8a.png}
\hspace*{-2em}
\includegraphics[width=.5\textwidth]{fig8b.png}
\vspace*{-2em}
\caption{$\sigma_{H_1}(ggF)/\sigma_{H^{SM}}(ggF)\times BR(H_1\to A_1
A_1)$ as function of $M_{H_1}$ (left) and $M_{A_1}$ (right). The color
coding is as in Fig. \ref{fig:1}.}
\label{fig:8}
\end{figure}
The dominant decay branching fractions of $A_1$ are very similar to the
ones of a SM-like Higgs boson of the same mass, i.e. dominantly into
$b\bar{b}$ and $\tau^+\tau^-$ if kinematically allowed. These
unconventional channels $H \to A_1 A_1 \to ...$ have been searched for
at LEP by OPAL~\cite{Abbiendi:2002qp,Abbiendi:2004ww,Abbiendi:2002in},
DELPHI~\cite{Abdallah:2004wy} and ALEPH~\cite{Schael:2010aw}. The
corresponding constraints are taken into account in {\sf NMSSMTools},
and explain the absence of sizeable signal rates for $M_{H_1} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\;
80$~GeV. For $M_{H_1} \;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 86$~GeV and, simultaneously, $0.25~
\mathrm{GeV} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\;
M_{A_1} \;\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 3.55$~GeV, first LHC analyses by CMS~\cite{Chatrchyan:2012cg}
have lead to upper limits on the signal cross
section for $H \to A_1 A_1 \to 4\mu$ which exclude some of the points in
this range of $M_{A_1}$.
For heavier $A_1$ leading to dominant $b\bar{b}$ and/or $\tau^+\tau^-$
decays, analyses of possible signals are certainly more difficult. At
least we find that, for $M_{H_1} \;\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}\; 80$~GeV, production cross
sections times branching fractions can be relatively large without
violating present constraints, which should motivate future analyses of
these channels.
Concerning the signal rates of the SM-like Higgs boson $H_2$ we
remark that all values allowed by the 95\% confidence level contours
in~\cite{Belanger:2013xza} in the planes of Higgs production via
(gluon fusion and ttH) -- (vector boson fusion and associate
production with W/Z) for Higgs
decays into $\gamma\gamma$, ZZ+WW and $b\bar{b}+\tau^+\tau^-$ have been
found by our scan.
Also possible are decays of $H_2$ into pairs of light CP-even or CP-odd
states $H_1$ or $A_1$. They are limited by the SM-like signal rates of
$H_2$, but branching fractions of up to 40\% are still allowed.
\section{Conclusions and outlook}
In spite of the recent constraints on the mass and the signal rates at
the LHC on a SM-like Higgs boson, upper bounds on signal rates generated
by first generation squarks and gluinos and upper bounds on dark matter
-- nucleus cross sections we have seen that large ranges of the
parameter space of the NUH-NMSSM remain viable. Within this scenario,
bounds from squark/gluino searches dominate the lower bounds on
fine-tuning which remain, on the other hand, considerably smaller than
in the (NUHM-)MSSM and more constrained versions of the NMSSM.
The mass of the LSP is barely constrained, up to some ``holes'' around 30
and 50 GeV, and can possibly be below 1~GeV. Due to its possibly
dominant singlino component, its direct detection cross section can be
considerably smaller than the neutrino background, which makes it
compatible with all future null-results in direct (and actually also
indirect) dark matter searches.
We have not discussed all possible NUH-NMSSM-specific phenomena at
colliders, which would be beyond the scope of the present paper. Here we
focussed on the properties of an additional lighter NMSSM-specific Higgs
boson $H_1$, in particular on its signal rates in channels which are
accessible at the LHC. These include the potentially promising diphoton
decay channel, but also $H_1$-decays into a pair of even lighter CP-odd
bosons $A_1$. Albeit taking into account all present constraints on
additional lighter Higgs bosons, wide ranges of $H_1$ and $A_1$ masses
remain to be explored.
Amongst additional NUH-NMSSM-specific phenomena at colliders -- induced
by a sing\-lino-like LSP and/or additional Higgs states -- are possibly
unconventional cascade decays of charginos and top- and bottom-squarks,
which require additional studies. Future work will also be dedicated to
the possibilities for and signatures of Higgs-to-Higgs decay cascades
induced by heavier Higgs states in the NUH-NMSSM.
\section*{Acknowledgements}
UE acknowledges support from the ERC advanced grant Higgs@LHC and from
the European Union Initial Training Networks INVISIBLES
(PITN-GA-2011-289442) and Higgs\-Tools (PITN-GA-2012-316704).
The authors acknowledge the support of France Grilles for providing
computing resources on the French National Grid Infrastructure.
\newpage
|
\section{Introduction}
\label{intro}
A merger between two galaxies, each with its own central supermassive black hole (SMBH), brings the two SMBHs to the center of the resultant merger-remnant galaxy. The pair is known as dual SMBHs when the black holes are separated by kiloparsec (kpc) scales, before the pair evolves into a gravitationally-bound SMBH binary and ultimately coalesces.
Since galaxy mergers can trigger gas inflows that fuel active galactic nuclei (AGNs; \citealt{SU98.1,CA01.2,TR12.1}) and the AGN fraction increases as the SMBH separation decreases from $\sim80$ kpc to 5 kpc \citep{EL11.1,KO12.1}, some of these dual SMBHs should be active. When one or both SMBHs power AGNs, the systems are known as offset AGNs and dual AGNs, respectively. These offset and dual AGNs are valuable for studies of galaxy evolution, since they are direct observational tracers of SMBH mass growth via gas accretion during mergers.
Dual AGNs have been popular targets of recent study (e.g., \citealt{CO09.3,BA12.1,KO12.1,TE12.1,VA12.1,BL13.1,IM14.1,WO14.1}), with the first systematic searches beginning in the last few years. Most of these searches focus on AGN spectra with double-peaked narrow emission lines (e.g., \citealt{CO09.1,WA09.1,XU09.1,LI10.1,SM10.1,GE12.1,BA13.1,CO13.1}), which can be produced by the relative motions of two narrow-line regions (NLRs) accompanying two AGNs moving in the host galaxy potential.
Double-peaked narrow AGN emission lines can can also be caused by disk rotation and the NLR structure of biconical AGN outflows (e.g., \citealt{HE81.1,CR00.1,VE01.1,WH04.1,DA05.1,CR10.1,FI11.1}). To determine the true nature of double-peaked AGNs there have been many follow-up observations, including optical spectroscopy, near infrared imaging, radio observations, {\it Hubble Space Telescope} imaging, and {\it Chandra} observations \citep{LI10.2, CO11.2, FU11.1,FU11.3,GR11.1,MC11.1,RO11.1,SH11.1,TI11.1,CO12.1,FU12.1,GR12.2,LI13.1}. These follow-up observations have led to several confirmations of double-peaked emission lines that are produced by dual AGNs \citep{FU11.3,LI13.1}.
In contrast, offset AGN candidates have so far received little attention despite their potential for yielding many more dual SMBH discoveries. Study of offset AGNs will also open the door to comparisons between offset AGN and dual AGN populations, which will uncover the details of AGN fueling during mergers and differing circumstances for fueling of one versus both SMBHs in a merger.
Offset AGNs have been observed, as in the example of the $z=0.0271$ disturbed disk galaxy NGC 3341 \citep{BA08.1,BI13.1}. This offset AGN was a serendipitous discovery, and it is a Seyfert 2 at a projected separation of $9\farcs5$ (5.2 kpc) from the nucleus of the host galaxy. The offset AGN also has a line-of-sight velocity that is blueshifted by 200 km s$^{-1}$ relative to the host galaxy nucleus. Several studies of AGNs in galaxy pairs, where pairs are defined by maximum transverse separations varying from 30 to 80 kpc and maximum line-of-sight velocity separations varying from 200 to 500 km s$^{-1}$, may have also detected offset AGNs, although they are not discussed specifically \citep{AL07.1,WO07.2,RO09.1,KO10.1,EL11.1}.
We conduct a systematic search for offset AGNs in the Sloan Digital Sky Survey (SDSS; \citealt{YO00.2}) via the spectral signature of AGN emission lines that are offset in line-of-sight velocity from systemic, as in NGC 3341. This approach is analogous to the technique of searching for dual AGNs via double-peaked narrow AGN emission lines. Similar to the case for double-peaked AGN emission lines, single-peaked AGN emission lines with velocity offsets can be produced by disk rotation, AGN outflows, recoiling SMBHs, dust obscuration, or dual SMBHs. Consequently, the velocity-offset AGNs we find here are candidates for offset AGNs, but confirmation of their true natures requires additional follow-up observations.
\begin{deluxetable*}{lllll}
\tabletypesize{\scriptsize}
\tablewidth{0pt}
\tablecolumns{5}
\tablecaption{Summary of Offset AGN Candidates}
\tablehead{
\colhead{SDSS Designation} &
\colhead{Host Galaxy Redshift} &
\colhead{$\Delta v_\mathrm{Balmer}$ (km s$^{-1}$)} &
\colhead{$\Delta v_\mathrm{forbidden}$ (km s$^{-1}$)} &
\colhead{$\Delta v_\mathrm{weighted}$ (km s$^{-1}$)}
}
\startdata
SDSS J001828.09$-$003412.3 & $0.069291 \pm 0.000017$ & $-64.4 \pm 14.7$ & $-70.0 \pm 14.7$ & $-67.2 \pm 10.4$ \\
SDSS J002312.35+003956.2 & $0.072648 \pm 0.000020$ & \phn \phd $47.9 \pm 14.8$ & \phn \phd $54.8 \pm 14.7$ & \phn \phd $51.4 \pm 10.4$ \\
SDSS J003908.37$-$105833.0 & $0.065406 \pm 0.000020$ & $-72.6 \pm 14.8$ & $-52.1 \pm 14.8$ & $-62.4 \pm 10.5$ \\
SDSS J003948.38$-$090834.5 & $0.037475 \pm 0.000023$ & \phn \phd $52.6 \pm 14.7$ & \phn \phd $49.7 \pm 14.7$ & \phn \phd $51.1 \pm 10.4$
\enddata
\tablecomments{(This table is available in its entirety in a machine-readable form in the online journal. A portion is shown here for guidance regarding its form and content.)}
\label{tbl-1}
\end{deluxetable*}
This work is the third systematic search for offset AGNs. The first two searches identified offset AGN candidates via velocity-offset AGN emission lines in the DEEP2 Galaxy Redshift Survey \citep{CO09.1} and in the AGN and Galaxy Evolution Survey (AGES; \citealt{CO13.1}). These searches uncovered 30 offset AGN candidates at a mean redshift $\bar{z}=0.7$ and 5 offset AGN candidates at $\bar{z}=0.25$, respectively. The search for offset AGN candidates in SDSS has the potential to yield many more candidates, given the much larger size of the SDSS spectroscopic catalog compared to those of DEEP2 and AGES. Furthermore, the search through SDSS ($\bar{z}=0.1$) will fill in a population of offset AGN candidates at low redshifts, which will enable follow-up observations to resolve offset AGNs with $<$ kpc projected offsets.
We note that while an offset AGN is the case of dual SMBHs where one SMBH is active and the other is quiescent, our search here focuses by necessity on {\it detectable} offset AGNs. Detectable offset AGNs are dual SMBH systems where one SMBH is an AGN and the other SMBH is not detected as an AGN, either because it is a quiescent SMBH or an AGN that is obscured by dust or confused with star formation.
We assume a Hubble constant $H_0 =70$ km s$^{-1}$ Mpc$^{-1}$, $\Omega_m=0.3$, and $\Omega_\Lambda=0.7$ throughout, and all distances are given in physical (not comoving) units.
\section{Selecting Offset AGN Candidates}
\label{selection}
We begin with the SDSS DR7 catalog of $z<0.21$ objects identified as galaxies by the SDSS pipeline \citep{AB09.1} and the OSSY catalog \citep{OH11.1} of velocity dispersion, line position, and flux measurements for SDSS DR7 spectra. OSSY uses codes for penalized pixel-fitting ({\tt pPXF}) and gas and absorption line fitting ({\tt gandalf}; \citealt{SA06.1}) to
simultaneously fit an entire spectrum using stellar templates for the stellar kinematics and Gaussian templates for the emission components. This process returns high quality measurements of wavelengths, fluxes, and widths of absorption and emission features for each SDSS spectrum. OSSY defines the quality of their fits to the spectra using the level of the formal uncertainties in the flux densities, which is the statistical noise, and the level of fluctuations in the fit residuals, which is the residual noise.
From the catalog of $z<0.21$ galaxies in SDSS DR7, we select the 20098 spectra identified as Type 2 AGNs in \cite{BR04.1}, which requires that the \mbox{[\ion{O}{3}] $\lambda$5007}, \mbox{H$\beta$}, \mbox{H$\alpha$}, and \mbox{[\ion{N}{2}] $\lambda$6584} $\,$ lines have signal-to-noise ratios greater than 3 and that the emission line flux ratios lie above the theoretically derived boundary between composite systems and AGNs \citep{KE01.2}.
We restrict the sample to those AGNs with robust fits to the absorption and emission line systems in the SDSS spectra. To do this, we examine the plot of residual-noise-to-statistical-noise ratio against median signal-to-statistical-noise ratio for the OSSY catalog. We define quality fits as those that deviate by less than $3\sigma$ from the median line in this plot (see \citealt{OH11.1}). This criterion reduces the sample to 18314 AGNs, which we define to be our ``parent sample" of AGNs from which we will identify offset AGN candidates.
Our purpose is to select the AGN spectra with kinematic signatures of offset AGNs. Specifically, we search for line-of-sight velocity offsets of the AGN-fueled emission lines relative to the stellar absorption features, because such velocity offsets are an expected consequence of the bulk motion of an AGN brought into a merger-remnant galaxy. Since other effects such as AGN outflows, disk rotation, gravitational recoil of SMBHs, and dust obscuration are well known to produce velocity offsets in AGN emission lines, we carefully construct criteria that will select {\it for} offset AGNs and {\it against} these other kinematic effects.
Our selection criteria for offset AGNs are as follows. In the OSSY catalog, all forbidden lines are forced to have the same kinematics, while all the Balmer lines are fit with a separate kinematical model (e.g., \citealt{TR04.2}). We use the velocities of the forbidden lines, Balmer lines, and stellar absorption features measured in OSSY to derive line-of-sight velocity offsets of the forbidden and Balmer lines relative to the stars.
First, we require that the line-of-sight velocity offsets of the forbidden lines and of the Balmer lines agree to $1\sigma$. This criterion aids in separating the offset-AGN candidates (where all emission lines should have consistent velocity offsets, due to the bulk motion of the AGN) from the AGN outflows (where there may be a stratified velocity structure, with different velocity offsets in the Balmer and forbidden lines; e.g., \citealt{ZA02.1}) and the gravitationally recoiling SMBHs (where velocity offsets are seen in the broad emission lines; e.g., \citealt{JU13.1}). We measure the line-of-sight velocity offsets relative to the measured redshift of the stellar absorption lines, which we take to be systemic. In addition to the uncertainties on the line velocities reported in OSSY, we also add a systematic uncertainty based on the variations in multiple SDSS observations of the same system. In the 1289 AGNs in our sample that were observed more than once with SDSS, we find a mean redshift difference of 10 km s$^{-1}$ between two epochs of observations. Consequently, we add a 10 km s$^{-1}$ systematic uncertainty to each velocity. We find that 15173 AGNs have forbidden lines and Balmer lines whose line-of-sight velocity offsets are consistent to $1\sigma$.
Second, we select the AGNs with line-of-sight velocity offsets that are greater than $3\sigma$ in significance. This criterion is designed to select against stationary AGN and SDSS spectra that may have been taken with miscentered fibers (which could result in rotating gas producing a slight redshifted or blueshifted velocity offset in the emission lines), but as a side effect it also removes bona fide offset AGNs that have small line-of-sight velocity offsets due to projection effects (we account for this selection effect in Section~\ref{small}). After this cut 544 AGNs remain, and we define them as the ``velocity-shifted sample" of AGNs.
Third, we require that the AGN emission line profiles are symmetric. The goal here is to select against AGN outflows, whose commonly asymmetric line profiles have been well documented (e.g., \citealt{HE81.1,WH85.1,CR00.2,TA01.1,ZA02.1,GR05.1,DA06.1,KO07.2,WA11.1}). To do this, we fit the spectra with {\tt gandalf} to obtain the continuum-subtracted emission line profiles. Then, we borrow the approach used in studies of high redshift galaxies to distinguish asymmetric Ly$\alpha$ emission lines at high redshift from the symmetric emission lines, such as H$\alpha$, of lower redshift interlopers (e.g., \citealt{DI07.1}). Following \cite{KA06.1}, we define the weighted skewness as the statistical skewness of a line's continuum-subtracted flux, over the wavelength range where the continuum-subtracted flux values are greater than $10\%$ of the line's peak continuum-subtracted flux value. Since the blended \mbox{[\ion{N}{2}]} and \mbox{H$\alpha$} $\,$ lines prevent accurate measurements of their skewnesses, we focus on the \mbox{H$\beta$} $\,$ and \mbox{[\ion{O}{3}]} $\,$ lines. We require weighted skewnesses less than 0.5 \citep{BU79.1} in the continuum-subtracted \mbox{H$\beta$} $\,$ and \mbox{[\ion{O}{3}]} $\,$ emission lines, and these criteria are met by 365 AGNs.
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig1.eps}
\end{center}
\caption{Histograms of line-of-sight velocity offsets in the SDSS $z<0.21$ sample of AGNs. For the parent sample of 18314 AGNs, we show the velocity offsets of the forbidden emission lines (red dotted) and the Balmer emission lines (blue dashed) relative to the stellar absorption features. The black histogram illustrates the weighted velocity offsets of the 351 offset AGN candidates.}
\label{fig:vhist}
\end{figure}
\begin{figure*}
\hspace*{.4in}
\includegraphics[width=.6\linewidth]{fig2a.eps}
\includegraphics[width=.275\linewidth]{fig2b.eps} \\
\hspace*{.4in}
\includegraphics[width=.6\linewidth]{fig2c.eps}
\includegraphics[width=.275\linewidth]{fig2d.eps} \\
\hspace*{.4in}
\includegraphics[width=.6\linewidth]{fig2e.eps}
\includegraphics[width=.275\linewidth]{fig2f.eps} \\
\hspace*{.4in}
\includegraphics[width=.6\linewidth]{fig2g.eps}
\includegraphics[width=.275\linewidth]{fig2h.eps}
\caption{\scriptsize{SDSS spectra and imaging of four example offset AGN candidates. Left: the SDSS spectra are plotted in the restframe of each galaxy, based on the redshift of the stellar absorption features (such as Ca H+K, whose rest wavelengths are shown by the dotted vertical lines). Middle: SDSS spectra illustrating some of the AGN-fueled emission lines, plotted in the restframe of the galaxy's stars. The wavelengths of \mbox{H$\beta$}, [O III] $\lambda$4959, and [O III] $\lambda$5007, in the restframe of the galaxy's stars, are shown as dotted vertical lines, and the velocity shifts $\Delta v$ in the emission lines relative to the galaxy restframe are given. Right: $50^{\prime\prime} \times 50^{\prime\prime}$ SDSS $gri$ color-composite images of the galaxies, with $5^{\prime\prime}$ scale bars shown.}}
\label{fig:examples}
\end{figure*}
Finally, we remove the AGNs with known double-peaked emission lines, which are produced by rotating gas disks, biconical outflows, or dual AGNs (e.g., \citealt{VE01.1,WH04.1,CO09.1,CR09.1,XU09.1,RO10.1,FI11.1,GR11.2}). After removing the objects that were identified as double-peaked AGNs in searches through SDSS spectra \citep{WA09.1,LI10.1,SM10.1,GE12.1}, the result is 351 AGNs. We measure the weighted velocity offset for each AGN as the mean of the Balmer and forbidden line-of-sight velocity offsets, weighted by their inverse variances, and the weighted velocity offsets are shown in Table~\ref{tbl-1} and Figure~\ref{fig:vhist}.
These 351 AGNs are our offset AGN candidates, and examples of the offset AGN candidates are shown in Figure~\ref{fig:examples}. We focus the rest of the paper on analysis of the offset AGN candidates and comparisons to related groups of AGNs.
\section{Fraction of AGNs that \\ Are Offset AGNs}
\label{small}
Since we are sensitive to only the projection of an AGN's velocity along the line of sight, our selection of offset AGN candidates is incomplete. For instance, our criterion that offset AGN candidates have velocity offsets that are greater than 3$\sigma$ in significance excludes AGNs that have small, but real, projected velocity offsets. Here, we estimate the fraction of all AGNs that are offset AGNs.
First, we assume that a fraction $f_\mathrm{offset}$ of all AGNs in our parent sample are in fact offset AGNs and that every active galaxy is equally likely to host an offset AGN. Then, we assume that an offset AGN is orbiting in the potential of the host galaxy and that the three-dimensional velocity of an offset AGN is given by the three-dimensional host galaxy velocity dispersion. This is a reasonable assumption, since the velocity dispersion can be used as a proxy for a galaxy's gravitational potential, via the Jeans equations \citep{JE15.1}. In this case, the observed line-of-sight velocity is $v_{obs}= \sigma_* \cos\theta$, where $\sigma_*$ is the line-of-sight velocity dispersion and $\theta$ is the polar angle of the observer to the AGN in a spherical coordinate system. Next, we assume that the AGN's velocity has a random orientation in the plane of the host galaxy, so that $|\cos\theta|$ has a random value between 0 and 1.
Then, we iterate through values of $f_\mathrm{offset}$ and compare the predicted distribution of velocity offsets to the distribution of actual observed velocity offsets for the offset AGN candidates. For each value of $f_\mathrm{offset}$, we use Monte Carlo realizations to draw 1000 distributions of predicted observed velocity offsets. For each predicted distribution, we measure the chi-squared difference between the predicted and observed velocity offsets. We compare the velocity offsets that have absolute values greater than 50 km s$^{-1}$. We then take the median value of the 1000 chi-squared measurements at each $f_\mathrm{offset}$ value, and we find that $f_\mathrm{offset}=0.05$ is the best fit to the observed velocity offsets (Figure~\ref{fig:lowv}).
We also extend this approach to include AGNs that would have been classified as offset AGN candidates with $|v_{em}-v_{abs}|>50$ km s$^{-1}$, except that their line-of-sight velocity offsets are not greater than 3$\sigma$ in significance. This adds 196 AGNs and provides an upper limit to the number of offset AGN candidates with $|v_{em}-v_{abs}|>50$ km s$^{-1}$ in our sample. When we vary $f_\mathrm{offset}$ and match to the data, as described above, we find that $f_\mathrm{offset}=0.08$ provides the best fit to this sample (Figure~\ref{fig:lowv}).
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig3.eps}
\end{center}
\caption{Histograms of the velocity offsets of the offset AGN candidates (solid histograms) and expectations of the velocity offset distributions drawn from the AGN parent sample (dashed histograms). The 351 offset AGN candidates are shown in black, while the red histogram includes the addition of 196 AGNs with $|v_{em}-v_{abs}|>50$ km s$^{-1}$ and that only missed offset candidate classification because their velocity errors are too large. The black (red) dashed histograms show the expected velocity offset distributions drawn from Monte Carlo realizations, assuming that 5\% (8\%) of the parent AGN sample are offset AGNs that have random orientations in the host galaxy planes.}
\label{fig:lowv}
\end{figure}
As a test of the robustness of these figures, we also determine $f_\mathrm{offset}$ using the subsample of offset AGN candidates with velocity offsets that have absolute values greater than 70 km s$^{-1}$. For this sample of 189 offset AGN candidates, we find $f_\mathrm{offset}=0.04$. When we add the 18 AGNs in this velocity range that only miss offset AGN classification because their line-of-sight velocity offsets are not greater than 3$\sigma$ in significance, we find the same result. We use this as a lower limit on the value of $f_\mathrm{offset}$.
Consequently, we estimate that 4\% -- 8\% of AGNs in our sample could in fact be offset AGNs. With these estimates we find that our selection technique, and specifically the requirement that the velocity offsets are greater than 3$\sigma$ in significance, could be missing $\sim300 $ -- 700 offset AGNs with offset velocities $|v_{em}-v_{abs}|<50$ km s$^{-1}$.
\section{Nature of the Offset AGN Candidates}
\label{nature}
We shaped our selection criteria to select the best candidates for offset AGNs and to avoid offset emission lines produced by recoiling SMBHs or by rotating gas disks or AGN outflows, which may also conspire with dust obscuration. Here we examine in detail whether the offset AGN candidates are consistent with being produced by recoiling SMBHs, rotating disks, AGN outflows, or dust obscuration.
\subsection{Recoiling SMBHs}
Velocity offset emission lines can also be produced by a recoiling SMBH, where the recoil is the result of gravitational wave emission after the merger of two SMBHs at the center of a merger-remnant galaxy. However, such recoiling SMBHs, as well as subparsec-scale binary SMBHs, produce velocity offsets in the {\it broad} AGN emission lines and not the narrow AGN emission lines (e.g., \citealt{GA84.1,BO07.1,ER12.1,JU13.1,SH13.1}). Via our criterion that the velocity offsets of the forbidden lines and the Balmer lines agree to within $1\sigma$, we have removed velocity offsets that appear in the broad lines only.
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig4.eps}
\end{center}
\caption{The line-of-sight velocity offsets plotted against the weighted skewnesses of the \mbox{H$\beta$} $\,$ (left) and \mbox{[\ion{O}{3}] $\lambda$5007} $\,$ (right) emission lines. The black points represent the velocity-shifted sample of 544 AGNs and the points circled in green represent the 351 offset AGN candidates, which are the subset of the 544 AGNs that were selected for their line symmetries and their lack of double peaks. Negative (positive) velocity offset signals a blueshifted (redshifted) emission line, while negative (positive) skewness indicates that a line is skewed to the blue (red). In both \mbox{H$\beta$} $\,$ and \mbox{[\ion{O}{3}] $\lambda$5007}, there are only mild correlations between blueshifted (redshifted) velocity offsets and blue (red) skewnesses. }
\label{fig:skew}
\end{figure}
\newpage
\subsection{Rotating Disks}
Some of the velocity-offset emission lines in our sample may be caused by rotating disks, where the disk gas may be clumpy or partially obscured. If we see only one side of the disk, then we may expect a correlation between the velocity offsets and the line asymmetries. That is, we might expect the line to be asymmetric in the same direction of the velocity offset, with redshifted velocity offsets correlated to positive skewness and blueshifted velocity offsets correlated to negative skewness. For example, if we observe only the blueshifted side of the disk then we would observe an overall blueshift in the emission lines relative to systemic and more blue light relative to red light, which produces a negative skewness in the observed emission lines.
To test the significance of such rotating disks as a contaminant in our sample, we compare the 351 offset AGN candidates to the velocity-shifted sample of 544 AGNs (Figure~\ref{fig:skew}). We use the skewnesses as measured in Section~\ref{selection}. For \mbox{H$\beta$}, we find mild correlations between $v_{em}-v_{abs}$ and \mbox{H$\beta$} $\,$ skewness for the velocity-shifted AGN sample (the Spearman's rank correlation coefficient is 0.13 and the significance level of its deviation from zero is 0.002) and for the offset AGN candidates (Spearman's rank correlation coefficient 0.11, significance level 0.05). For \mbox{[\ion{O}{3}] $\lambda$5007}, we also find mild correlations between $v_{em}-v_{abs}$ and \mbox{[\ion{O}{3}] $\lambda$5007} $\,$ skewness for the velocity-shifted AGN sample (Spearman's rank correlation coefficient 0.06, significance level 0.14) and for the offset AGN candidates (Spearman's rank correlation coefficient 0.12, significance level 0.03).
Since we find no significant correlation of velocity offset with skewness, we conclude that our sample of offset AGN candidates likely is not contaminated by many rotating disks.
\subsection{AGN Outflows}
Comparisons of the velocity offset of the narrow AGN emission line to the velocity dispersion of the host galaxy can also shed light on the source of the velocity offset. The motion of offset AGNs is dominated by dynamical friction from the host galaxy stars, so offset AGNs should follow the potential of the host galaxy. Due to projection effects, the measured line-of-sight velocity offset of an offset AGN should be less than or equal to the velocity dispersion. In contrast, the measured line-of-sight velocity offset of an AGN outflow can be less than, equal to, or greater than the velocity dispersion (e.g., \citealt{NE08.1,CR10.1,TO10.1,LI13.2}).
To learn about the nature of the SDSS offset AGN candidates, we compare the absolute values of the velocity offsets to the velocity dispersions in Figure~\ref{fig:vdisp}. For the parent population of 18314 AGNs, 98\% (99\%) of the velocity offsets of the forbidden (Balmer) emission lines fall below the velocity dispersions. For the 351 offset AGN candidates, 97\% of the weighted velocity offsets fall below the velocity dispersions.
There are six offset AGN candidates with weighted velocity offsets that are $>1\sigma$ above the velocity dispersions. These may be turbulent merging galaxies where the velocity dispersion is changing rapidly (e.g., \citealt{JO09.1,ST12.1}), or they may be examples of AGN outflow interlopers. Follow-up observations are necessary to determine their true natures (Section~\ref{conclusions}).
The relative numbers of redshifted and blueshifted AGN emission lines also offer clues into the sources of the velocity offsets. If offset AGNs have random velocity projections to the line of sight, then there should be equal numbers of offset AGNs with observed redshifts and offset AGNs with observed blueshifts. In contrast, AGN outflows are known to preferentially result in observations of blueshifted emission lines, since redshifted lines are often obscured by the AGN torus (e.g., \citealt{ZA02.1}). Among our offset AGN candidates, 170 ($48^{+3}_{-2}\%$) exhibit redshifted lines and 181 ($52^{+2}_{-3}\%$) exhibit blueshifted lines, which is consistent with the equal numbers of redshifted and blueshifted observations expected for offset AGNs.
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig5.eps}
\end{center}
\caption{Absolute values of line-of-sight velocity offsets plotted against stellar velocity dispersions. For clarity, we show a random subsample of 10\% of the parent sample of 18314 AGNs, where the velocity offsets of the forbidden emission lines are shown in red and the velocity offsets of the Balmer emission lines are shown in blue. The black points illustrate the weighted velocity offsets of the 351 offset AGN candidates. The dashed line shows a 1:1 correlation between line-of-sight velocity offset and stellar velocity dispersion.}
\label{fig:vdisp}
\end{figure}
\begin{figure*}
\begin{center}
\includegraphics[width=18cm]{fig6.eps}
\end{center}
\caption{Histograms of the distributions of the parent sample of 18314 AGNs (black) and the 351 offset AGN candidates (red) in redshift (left), physical size of the SDSS fiber (middle), and velocity dispersion (right). For each distribution the median value is marked with a dotted vertical line.}
\label{fig:z_kpc_vdisp}
\end{figure*}
\subsection{Dust Obscuration}
If dust obscures one of the AGN in a dual AGN system, this could lead us to detect an offset AGN. Or, dust could conspire with disk rotation or AGN outflows to create the velocity offsets that we measure, by obscuring only the redshifted or blueshifted component of emission from a disk or outflow. We note that dust on small scales around the AGN torus should not affect our sample, which consists of Type 2 AGNs. In general the amount of larger scale dust obscuration increases with a galaxy's inclination, with the most obscuration occurring for edge-on galaxies, which have inclinations of 90$^\circ$.
As a test of whether dust is a large contaminant in our offset AGN candidates, we measure the inclinations of their host galaxies and compare to the inclinations of the host galaxies of the AGN parent population. We measure a galaxy's inclination via the exponential fit to the ratio of semiminor to semimajor axis $b/a$ in $r$ band, given by {\tt expAB\_r} with error {\tt expABErr\_r}, in the SDSS PhotObj catalog. The median error on the inclination for the AGN sample is $0.7^\circ$. We find that the offset AGN candidates have typical inclinations (mean inclination $51^\circ$, with a standard deviation of $14^\circ$) that are consistent with those of the AGN parent sample (mean inclination $48^\circ$, with a standard deviation of $15^\circ$).
We also test the dust obscuration hypothesis by examining the AGN luminosities as a function of host galaxy inclinations. If dust plays a significant role in our offset AGN sample, then we expect more obscured (fainter) AGNs as inclination increases. Instead, we find that the AGN bolometric luminosities are independent of inclination (Spearman's rank correlation coefficient 0.06, significance level 0.26).
Based on the distributions of inclinations and AGN luminosities, we find no evidence for dust obscuration as a large contributor to the velocity offsets we measure in the offset AGN candidates.
As a more thorough test of the role of dust obscuration in producing velocity-offset AGN emission features, we suggest that a large, complete sample of integral field spectrograph observations of active galaxies would be useful. Artificial dust reddening could be added to the data in different spatial locations on the galaxy in a systematic fashion, as a test of the effect on the observed stellar absorption and AGN emission kinematics.
\section{Results}
\subsection{Comparison to Parent AGN Population}
The offset AGN candidates reflect the parent AGN population in overall distributions of redshift, physical size of the SDSS fiber, and stellar velocity dispersion, but the median values differ (Figure~\ref{fig:z_kpc_vdisp}).
The median stellar velocity dispersion in the offset AGN candidates' host galaxies ($\sigma_*=161$ km s$^{-1}$) is higher than the median stellar velocity dispersion of the parent population ($\sigma_*=123$ km s$^{-1}$). This could be explained if the offset AGN sample consists of galaxy mergers that exhibit high velocity dispersions because the stars are not yet dynamically relaxed. Or, the higher stellar velocity dispersions could reflect a selection effect from our approach to identifying offset AGN candidates. We select offset AGN candidates as AGNs that have velocity offsets that are greater than $3\sigma$ in significance, which eliminates AGNs with velocity offsets $\lesssim 50$ km s$^{-1}$. If we assume that offset AGNs follow the orbital velocity of the host galaxy stars and that the offset AGNs' orbits are randomly oriented in the host galaxy planes (Section~\ref{small}), then our sample of offset AGN candidates is biased towards galaxies with higher velocity dispersions.
We also find that the offset AGN candidates have somewhat higher redshifts (median $z=0.104$) than the parent AGN population (median $z=0.085$) and, correspondingly, the physical size of the SDSS fiber is somewhat larger on average (a median of 5.7 h$_{70}^{-1}$ kpc for the offset candidates, compared to a median of 4.8 h$_{70}^{-1}$ kpc for the parent population). This may reflect the bias towards higher $\sigma_*$ for the offset AGN candidates, as discussed above, if $\sigma_*$ correlates with redshift in the SDSS catalog (e.g., \citealt{TH13.1}). Further, if the offset AGN catalog contains a higher fraction of galaxy mergers, the higher median redshift may reflect the increasing incidence of galaxy mergers with redshift (e.g., \citealt{CO03.2,LI08.1,LO11.1}).
\begin{figure*}
\begin{center}
\includegraphics[width=18cm]{fig7.eps}
\end{center}
\caption{Histograms of ${f}_\mathrm{vote}$, the fraction of Galaxy Zoo votes for a morphology classification for the offset AGN candidates (dotted histograms) and the mean fraction of votes for the same morphology classification for a comparison sample of AGNs matched in color, stellar mass, and redshift (solid histograms). The vote fractions for spiral, elliptical, and merger classifications are shown.}
\label{fig:gz}
\end{figure*}
\subsection{Morphologies and Environment}
If the offset AGN candidates are in fact offset AGNs in ongoing galaxy mergers, then we expect the candidates to preferentially reside in host galaxies that have merger morphologies or elliptical morphologies (since merger remnants have properties similar to ellipticals; e.g., \citealt{BA88.2,HE92.1}) and reside in dense environments with many other galaxies. Color and stellar mass are well known to correlate with morphology and environment.
Massive, red galaxies preferentially occur with elliptical morphologies, while less massive, blue galaxies preferentially occur with spiral morphologies (e.g., \citealt{DE61.1,ST01.2}). Furthermore, red galaxies are more clustered with other galaxies, while blue galaxies are less clustered with other galaxies (e.g., \citealt{BO92.2,YA05.2,CO06.3}). Finally, the occurrence rate of galaxy mergers has been observed to increase with redshift (e.g., \citealt{CO03.2,LI08.1,LO11.1}). We account for these underlying trends in morphology by comparing the offset AGN candidates to samples of SDSS AGNs that are matched in color, stellar mass, and redshift. Consequently, any remaining trends with morphology or environment can be attributed to the offset AGN candidates themselves.
For each offset AGN candidate, we build a comparison sample of AGNs (selected from the SDSS DR7 catalog of $z<0.21$ galaxies, as described in Section~\ref{selection}) that have the same $u-r$ color, stellar mass, and redshift, to within $10\%$, as the candidate. We construct these comparison samples for the 349 of the 351 offset AGN candidates that have existing stellar mass measurements based on fits to the photometry \citep{KA03.2,SA07.2}, which we obtain from the MPA-JHU SDSS DR7 data product\footnote{\url{http://www.mpa-garching.mpg.de/SDSS/}}. Colors are obtained from $u$ and $r$ magnitudes measured in the SDSS photometry pipeline. We choose the $10\%$ range because it produces comparison sample sizes of $\sim100$ for most offset AGN candidates; the median number of AGNs in a comparison sample is 89. There are 44 offset AGN candidates that have comparison groups with fewer than 10 members, and we do not include them in the morphology analysis.
Here, we examine the morphologies and environments of the remaining 305 offset AGN candidates that have comparison samples of at least 10 AGNs. We find a small overall preference (with large errors) for offset AGN candidates to occur in galaxies with elliptical, S0, and merger morphologies, as well as a bias for offset AGN candidates, on average, to occur in more clustered environments.
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig8.eps}
\end{center}
\caption{Histograms of $f_\mathrm{vote,offset \, AGN}-\bar{f}_\mathrm{vote,AGN}$, the differences between the fraction of Galaxy Zoo votes for a morphology classification for an offset AGN candidate $f_\mathrm{vote,offset \, AGN}$ and the mean fraction of Galaxy Zoo votes for the same morphology classification for a comparison sample of AGNs matched in color, stellar mass, and redshift $\bar{f}_\mathrm{vote,AGN}$. Fractions of the Galaxy Zoo votes for spiral (blue), elliptical (red), and merger (black) classifications are shown. For each distribution the mean value is marked with a dotted vertical line.}
\label{fig:gz_differences}
\end{figure}
\subsubsection{Morphology}
We examine the host galaxy morphologies of the offset AGN candidates using two different visual morphology indicators: one that is qualitative, and one that is more quantitative. The qualitative approach to visual morphologies is Galaxy Zoo, a catalog of visual morphology classifications assigned by citizen scientists to $\sim900,000$ galaxies in SDSS DR6 \citep{LI11.1}. Multiple citizen scientists vote to classify each galaxy as an elliptical, clockwise spiral, anti-clockwise spiral, edge on spiral galaxy, or merger. A known bias is that galaxies that are distant, small, or faint are more likely to be classified as ellipticals, even though they may be spiral galaxies with arms that are difficult to discern \citep{BA09.1}. A bias correction has been applied to account for this effect, based on the assumption that the fractions of galaxies with each morphology classification remains constant for a given galaxy size and luminosity. This correction resulted in a catalog of debiased vote fractions for the galaxies \citep{LI11.1}. We use the debiased vote fractions of elliptical, combined spiral (which combines the three types of spiral classifications), and merger for our analysis.
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig9.eps}
\end{center}
\caption{Histograms of the probabilities $P$ of a Bayesian automated morphology classification for the offset AGN candidates (dotted histograms) and the median probabilities of the same morphology classification for a comparison sample of AGNs matched in color, stellar mass, and redshift (solid histograms). The probabilities for elliptical, S0, Sab, and Scd classifications are shown.}
\label{fig:bayesian}
\end{figure}
For each morphology category, we compare the offset AGN candidate's vote fraction to the mean vote fraction of the same classification for the AGN comparison sample (Figure~\ref{fig:gz}). On average, the offset AGN candidates received a slightly higher fraction (with large errors) of votes for elliptical and merger morphologies and a lower fraction of votes for spiral morphologies (Figure~\ref{fig:gz_differences}). On average, the offset AGN candidates have $2 \pm 27\%$ ($3 \pm 14\%$) higher vote fractions for elliptical (merger) morphologies than their comparison AGN samples. Conversely, the offset AGN candidates have $6 \pm 29\%$ lower vote fractions, on average, for spiral morphologies than their comparison AGN samples.
We also assess morphology using a catalog of Bayesian automated morphology classifications for $\sim700,000$ galaxies in SDSS DR7 \citep{HU11.1}, and this catalog takes a more quantitative approach to morphology classifications. This technique, which uses a machine learning algorithm trained on visual classifications of morphologies, assigns each galaxy a probability of being an elliptical (E), S0, early-type spiral (Sab), or late-type spiral (Scd) galaxy.
For each morphology class, we compare the offset AGN candidate's probability to the median probability of the same classification for the AGN comparison sample (Figure~\ref{fig:bayesian}). We find that, on average, the offset AGN candidates have ($9 \pm 26\%$, $9 \pm 23\%$, $5 \pm 24\%$, $2 \pm 13\%$) higher probabilities of (E, S0, Sab, Scd) classifications (Figure~\ref{fig:bayesian_difference}). The spread in probabilities for each morphology category is large, but overall the offset AGN candidates have the greatest enhancement in probabilities for the elliptical and S0 classifications.
In summary, the two morphology catalogs we use yield similar results. We find enhanced probabilities of (2--9\%,9\%,3\%), with large error bars, for an offset AGN candidate to reside in a galaxy with an elliptical, S0, or merger morphology, respectively, when compared to AGNs with similar colors, stellar masses, and redshifts.
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig10.eps}
\end{center}
\caption{Histograms of $P_\mathrm{offset \, AGN}-\tilde{P}_\mathrm{AGN}$, the differences between the probabilities of a Bayesian automated morphology classification for an offset AGN candidate $P_\mathrm{offset \, AGN}$ and the median probability of the same morphology classification for a comparison sample of AGNs matched in color, stellar mass, and redshift $\tilde{P}_\mathrm{AGN}$. Probabilities for elliptical (red), S0 (black), Sab (blue), and Scd (green) classifications are shown. For each distribution the mean value is marked with a dotted vertical line.}
\label{fig:bayesian_difference}
\end{figure}
\subsubsection{Environment}
To study the global environments of the offset AGN candidates, we examine the number of galaxies $N_{group}$ in each candidate's group. We take group membership from an SDSS DR7 group catalog \citep{TI11.2,WE12.1}, which uses the halo-based group finding algorithm of \cite{YA05.1}. This catalog identifies groups only to $z=0.1$, due to the limitations in robust group determination at higher redshifts.
Because of the redshift limit of the group catalog, only 117 offset AGN candidates are in the group catalog with comparison samples of at least 10 AGNs matched in color, stellar mass, and redshift. Of these 117 offset AGN candidates, 101 have $N_{group} \leq 5$ (field), 14 have $5 < N_{group} \leq 50$ (group), and 2 have $N_{group} > 50$ (cluster).
When we examine the differences between $N_{group}$ for an offset AGN candidate and the mean $\bar{N}_{group}$ for its comparison sample (Figure~\ref{fig:environment}), we find that 104 offset AGN candidates reside in groups with the same number of members, within 1$\sigma$, as the mean. Of the cases where $N_{group}$ for an offset AGN candidate differs by $>1\sigma$ from $\bar{N}_{group}$ for its comparison sample, there are 13 offset AGN candidates in richer environments and no offset AGN candidates in poorer environments. This hints towards a tendency for offset AGN candidates to live in more clustered environments than other AGNs.
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig11.eps}
\end{center}
\caption{Histogram of $N_\mathrm{group,offset \, AGN}-\bar{N}_\mathrm{group,AGN}$, the differences between the number of group members for an offset AGN candidate $N_\mathrm{group,offset \, AGN}$ and the mean number of group members for its comparison sample of AGNs matched in color, stellar mass, and redshift $\bar{N}_{group,AGN}$. A positive value of $N_\mathrm{group,offset \, AGN}-\bar{N}_\mathrm{group,AGN}$ indicates that an offset AGN candidate resides in a richer environment than the average environment of its comparison sample.}
\label{fig:environment}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig12.eps}
\end{center}
\caption{Histograms of the bolometric luminosities of the AGN samples. The parent sample of 18314 AGNs is illustrated by the solid black histogram, while the 351 offset AGN candidates are illustrated by the red histogram. For each distribution the mean value is marked with a dotted vertical line.}
\label{fig:lhist}
\end{figure}
\subsection{Fraction of Offset AGN Candidates Increases with AGN Luminosity}
AGN activity can be powered by gas inflows driven by galaxy mergers (e.g., \citealt{HE84.1,SA88.2,SP05.2}) as well as stochastic accretion of gas (e.g., \citealt{NO83.1,MA99.3,HO06.1,LI08.3}), but the relative roles of these two mechanisms and the dependence on AGN luminosity is not clear. It has been suggested that galaxy mergers may play a more significant role in fueling higher-luminosity AGNs (e.g., \citealt{HO09.1,TR12.1}), while other studies indicate that there is no correlation of mergers with AGN luminosity (e.g., \citealt{VI14.1}).
Based on visual classification of mergers, \cite{TR12.1} find that the fraction of AGNs associated with mergers increases with the AGN bolometric luminosity, from $\sim4\%$ at $\log(L_\mathrm{bol})\sim43$ to $\sim70\%$ at $\log(L_\mathrm{bol})\sim46$. However, this result is based on identifying mergers via visual morphology classifications of the host galaxies. It is difficult to build a clean sample of AGNs in mergers with this approach, given that there are known limitations in using visual morphologies to identify mergers and these limitations can lead to significant numbers of false classifications (e.g., \citealt{LO11.1}).
Offset AGNs are uniquely well suited to building a clean sample of AGNs in mergers, since offset AGNs are by definition actively accreting supermassive black holes in ongoing galaxy mergers. We make the first attempt at using offset AGNs for studies of merger-triggered AGN activity, using our sample of offset AGN candidates. A final robust measurement will be possible once we have used follow-up observations to identify the true offset AGNs in the sample.
First, we determine the AGN bolometric luminosities by converting the OSSY measurements of \mbox{[\ion{O}{3}] $\lambda$5007} $\,$ fluxes to \mbox{[\ion{O}{3}] $\lambda$5007} $\,$ luminosities, then using the observed scaling relation of AGN bolometric luminosity with \mbox{[\ion{O}{3}] $\lambda$5007} $\,$ luminosity \citep{HE04.3}. Next, we compare the AGN bolometric luminosities of the offset AGN candidates to the parent AGN sample (Figure~\ref{fig:lhist}). The mean luminosity of the offset AGN candidates is $\sim0.3$ dex higher ($\log(L_\mathrm{bol})=44.2 \pm 0.7$) than that of the parent AGNs ($\log(L_\mathrm{bol})=43.9 \pm 0.7$), although they are consistent to within 1$\sigma$.
Since AGN bolometric luminosity may scale weakly with velocity dispersion (e.g., \citealt{WO02.1}), we bin the AGNs by host galaxy velocity dispersion before examining how the offset AGN candidate fraction depends on bolometric luminosity. We then bin by bolometric luminosity and find that the fraction of offset AGN candidates in the parent AGN sample increases with AGN bolometric luminosity for each range in velocity dispersion (Figure~\ref{fig:offset_bin}).
Finally, we stack the four velocity dispersion bins, using the mean of the offset AGN candidate fractions (in each bolometric luminosity bin) weighted by their inverse variances. We find that the fraction of offset AGN candidates in the parent AGN sample increases by an order of magnitude with AGN bolometric luminosity, from $0.7\%$ to $6\%$ over the bolometric luminosity range $43 < \log(L_\mathrm{bol})$ [erg s$^{-1}$] $< 46$ (Figure~\ref{fig:offset_stack}). The best-fit power law to the fraction of offset AGN candidates, $f$, is
\begin{equation}
f=\left( \frac{L_\mathrm{bol}} {2.1 \times 10^{49} \mathrm{\, erg \, s}^{-1}} \right)^{0.36} \; .
\end{equation}
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig13.eps}
\end{center}
\caption{The fraction of offset AGN candidates in the parent AGN sample, binned by velocity dispersion and then binned by bolometric luminosity. In the fourth panel, there are no parent AGNs in the bin $45.5 < \log(L_\mathrm{bol})$ [erg s$^{-1}$] $< 46$. The fraction of offset AGN candidates increases with $L_\mathrm{bol}$ and with $\sigma_*$.}
\label{fig:offset_bin}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm]{fig14.eps}
\end{center}
\caption{The fraction of offset AGN candidates in the parent AGN sample, binned by velocity dispersion and then stacked and binned by bolometric luminosity. The velocity dispersion binning accounts for the bias towards selecting offset AGN candidates in galaxies with higher velocity dispersions. The dotted line shows the best-fit power law, $f=(L_\mathrm{bol}/2.1 \times 10^{49} \mathrm{\, erg \, s}^{-1})^{0.36}$.}
\label{fig:offset_stack}
\end{figure}
The general trend of increasing fraction of AGNs in mergers with an increasing AGN bolometric luminosity is the same as seen in \cite{TR12.1} over the same bolometric luminosity range. \cite{TR12.1} also fit a power law to their $f$ -- $L_\mathrm{bol}$ relation, and their exponent is the same ($0.4$) but their amplitude is lower ([$3 \times 10^{46}$ erg s$^{-1}]^{-0.4}$). However, we note that the true amplitude of our relation is not well constrained due to known biases in our selection method (Section~\ref{small}).
If these candidates are in fact offset AGNs, then the observed trend is direct observational evidence that galaxy mergers preferentially trigger high-luminosity AGNs. However, if some of the offset AGN candidates are AGN outflows, then the relation may be explained in part by outflows that are correlated with higher bolometric luminosities (e.g., \citealt{HA12.1,KI13.1,TO13.1,CI14.1}). Forthcoming follow-up observations will confirm the bona fide offset AGNs and enable a robust measurement of any trend of increasing offset AGN fraction with AGN bolometric luminosity.
\section{Comparison to Other Samples of Offset AGN Candidates}
We now compare to higher redshift samples of offset AGN candidates that were selected in the AGES and DEEP2 surveys \citep{CO09.1,CO13.1} using similar criteria to those applied to SDSS here. The offset AGN candidates in both AGES and DEEP2 have line-of-sight velocity offsets of forbidden and Balmer lines that agree to $1\sigma$ and that are $>3\sigma$ in significance. In SDSS, we find 351 offset AGN candidates ($1.9^{+0.1}_{-0.1}\%$ of AGNs) with 50 km s$^{-1} < |v_{em}-v_{abs}| <$ 410 km s$^{-1}$ at a mean redshift $\bar{z}=0.1$. In AGES, there are 5 offset AGN candidates ($2.9^{+1.9}_{-0.8}\%$ of AGNs) with 100 km s$^{-1} < |v_{em}-v_{abs}| <$ 220 km s$^{-1}$ at a mean redshift $\bar{z}=0.25$. In DEEP2, there are 30 offset AGN candidates ($33^{+5}_{-5}\%$ of AGNs in red galaxies) with 40 km s$^{-1} < |v_{em}-v_{abs}| <$ 270 km s$^{-1}$ at a mean redshift $\bar{z}=0.7$.
After accounting for projection effects that remove small line-of-sight velocity offsets from the offset AGN candidate samples, we find that $4\% - 8\%$ of SDSS AGNs could be offset AGNs (Section~\ref{small}) while $44\% - 60\%$ of DEEP2 AGNs could be offset AGNs \citep{CO09.1}. One possibility for this increase is that effects such as outflows and dust obscuration, which can produce velocity-offset AGN emission lines (Section~\ref{nature}), may play a greater role in higher-redshift AGNs.
Before comparing the offset AGN candidates between these three surveys, we must account for differences in the surveys' spatial scales, spectral resolutions, and parent AGN selections. One possible bias in the comparisons is that the surveys are probing emission produced on different spatial scales. SDSS uses $3^{\prime\prime}$ fibers that correspond to 5.5 kpc at typical redshift $z=0.1$, AGES uses $1\farcs5$ fibers that correspond to 5.0 kpc at typical redshift $z=0.2$, and DEEP2 uses slits with 7$^{\prime\prime}$ average lengths that correspond to 50 kpc at typical redshift $z=0.6$ of the offset AGN candidates. The dual AGN candidates in SDSS and DEEP2 have typical spatial separations of $\sim1$ kpc \citep{CO09.1,CO12.1}, so if the offset AGN candidates are related systems (dual supermassive black holes, but where only one is an AGN) then their spatial separations should be $\lesssim1$ kpc. Consequently, the different spatial scales probed by the three surveys would not bias the relative numbers of offset AGN candidates found in each.
A more significant bias arises from the different spectral resolutions of the three surveys ($R\sim1800$ for SDSS, $R\sim1000$ for AGES, and $R\sim5000$ for DEEP2). The high spectral resolution of DEEP2 enables resolution of much smaller velocity offsets than is possible with SDSS or AGES. Further, the DEEP2 search for offset AGN candidates was limited to AGNs in red host galaxies. In order to compare offset AGN candidates across the redshift range of the SDSS and DEEP2 surveys, we select the offset AGN subsamples with velocity offsets $|v_{em}-v_{abs}|>46$ km s$^{-1}$ (the minimum velocity offset detected by SDSS, which has a lower spectral resolution than DEEP2) and that reside in host galaxies matched in color ($-0.6 < u-r < 5.8$) and rest-frame absolute magnitude ($-23.0 < \mathrm{M}_r < -18.3$) to the DEEP2 AGN sample.
When we match on minimum velocity offset, color, and rest-frame absolute magnitude in this way, we find that $1.9^{+0.1}_{-0.1}\%$ (345/18105) of AGNs are offset AGN candidates in SDSS and $32^{+6}_{-5}\%$ (29/91) of AGNs are offset AGN candidates in DEEP2. Hence, we find that the fraction of offset AGN candidates increases by a factor of $\sim16$ from $z=0.1$ to $z=0.7$. This trend can be understood if these candidates are indeed offset AGNs produced by galaxy mergers, since the galaxy merger fraction also increases with redshift (e.g., \citealt{CO03.2,LI08.1,LO11.1}).
\section{Conclusions}
\label{conclusions}
We have searched SDSS for an expected kinematic signature of offset AGNs: galaxy spectra that have AGN emission lines that are offset in line-of-sight velocity from the stellar absorption lines. Offset AGNs are produced when a merger between two galaxies brings two supermassive black holes into a merger remnant galaxy, and one of the supermassive black hole is fueled as an AGN. We find 351 offset AGN candidates out of a sample of 18314 AGNs at $z<0.21$, and their velocity offsets span 50 km s$^{-1} < |v_{em}-v_{abs}| <$ 410 km s$^{-1}$. We examine the characteristics of these offset AGN candidates and what they reveal about galaxy mergers, and our main results are summarized below.
1. By assuming that the AGN velocity offsets we measure reflect the host galaxy stellar velocity dispersions and are randomly oriented in the host galaxy planes, we estimate that $4\%$ -- $8\%$ of AGNs are offset AGNs.
2. We find that the offset AGN candidates, when compared to AGNs matched in color, stellar mass, and redshift, are more likely on average (with large error bars) to reside in host galaxies with elliptical or merger morphologies. We also uncover hints that offset AGN candidates reside in more clustered environments than the AGNs matched in color, stellar mass, and redshift. These host galaxy morphologies and denser environments are consistent with expectations for offset AGNs, which are produced by mergers.
3. The fraction of AGNs that are offset AGN candidates increases with AGN bolometric luminosity, from $0.7\%$ to $6\%$ over the luminosity range $43 < \log(L_\mathrm{bol})$ [erg s$^{-1}$] $< 46$. This suggests that higher luminosity AGNs are preferentially triggered by galaxy mergers and not stochastic accretion processes, and a robust measurement of this trend will be possible with future confirmations of offset AGNs.
4. By comparing and carefully matching to offset AGN candidates in galaxy surveys at higher redshifts, we find that the fraction of AGNs that are identified as offset AGN candidates increases from $1.9^{+0.1}_{-0.1}\%$ at $z=0.1$ to $32^{+6}_{-5}\%$ at $z=0.7$, a factor of $\sim16$ increase over that redshift range. This trend is suggestive of offset AGNs that are produced by galaxy mergers and so trace the galaxy merger rate, which exhibits similar increases with redshift over this range.
We developed our offset AGN candidate selection criteria carefully to select for velocity offsets produced by offset AGNs and against velocity offsets produced by gravitational recoil of a SMBH, an outflow from a single AGN, or a rotating gas disk around a single AGN. Multiple tests indicate that the candidates have the traits of offset AGNs and not of recoiling SMBHs, outflows, or disk rotation, but follow-up observations are required to confirm the true nature of the candidates. Spatial resolution of AGNs that are spatially offset from their host galaxy centers would provide direct confirmation of offset AGNs, and could be accomplished through spatially resolved spectroscopy, X-ray observations, or radio observations to pinpoint the spatial locations of the AGNs.
The selection of offset AGN candidates represents the first step towards building a catalog of confirmed offset AGNs, which would be the first of its kind and usher in a new angle of approach to observational studies of galaxy mergers and AGN fueling. Such studies are often limited by the difficulty of observationally identifying AGNs associated with ongoing galaxy mergers. A catalog of offset AGNs, built from candidates like those presented here, will enable unparalleled precision in resolving the link between galaxy mergers and AGN activity because offset AGNs are, by definition, actively accreting supermassive black holes in the midst of galaxy mergers.
\acknowledgements{We thank the referee for a prompt and useful report. We also gratefully acknowledge insightful discussions with Scott Barrows, Karl Gebhardt, Francisco M\"uller S\'anchez, and Kyuseok Oh.}
\newpage
\bibliographystyle{apj}
|
\section{Introduction}
The infrared structure of massless quantum electrodynamics in three dimensions (QED3) is an old problem~\cite{Pisarski:1984dj}.
Like other three-dimensional theories, QED3 is plagued by
infrared divergences.
It is widely believed that these signal a ground state that is far from the perturbative vacuum.
Two possibilities have been entertained~\cite{Appelquist:2004ib}.
One is that the infrared physics of the theory is described by confinement of
charge, with a concomitant dynamical mass for the fermions.
The other is that dynamical charges screen the Coulomb interaction, so that charges at large distances interact via a power law potential.
In the first case the scale of the theory's physics is presumably set by the coupling $e^2$, which has dimensions of mass.
In the second case, there is no scale in the infrared physics, but rather an
emergent conformal symmetry.
A number of studies have concluded that confinement and mass generation occur
if the number of (four-component) fermions, $N_f$, is small.
It is clear, on the other hand, that in the large-$N_f$ domain the theory is
screened and conformal.
Estimates of $N_c$, the critical value of $N_f$ that divides the two domains,
have varied widely in various analytical approaches \cite{Appelquist:1986fd,Appelquist:1988sr,Nash:1989xx,Pennington:1990bx,Curtis:1992gm,Appelquist:1994ui,Ebihara:1994wm,Maris:1996zg,Aitchison:1997ua,Gusynin:1998kz,Appelquist:1999hr,Gusynin:2003ww,Fischer:2004nq,Kaveh:2004qa,Bashir:2008fk,Goecke:2008zh,Braun:2014wja}
and lattice simulations \cite{Dagotto:1988id,Dagotto:1989td,Azcoiti:1993ey,Alexandre:2001pa,Hands:2002dv,Lee:2002id,Hands:2004bh,Fiore:2005ps,Strouthos:2008kc}
(for interim summaries see~\cite{Appelquist:2004ib,Strouthos:2008kc}).
We can explore this question via the renormalization group, guided by our knowledge of four-dimensional non-Abelian gauge theories.
The running of the coupling $e^2(q)$ is governed by the beta function,
\begin{equation}
\frac{de^2}{d\log q}=N_fb_1 e^4/q+\cdots,
\label{RG1}
\end{equation}
where $b_1>0$ is the one-loop coefficient.
Upon defining a dimensionless coupling $g^2(q)=e^2/q$, we have
\begin{equation}
\frac{dg^2}{d\log q}=-g^2+N_fb_1 g^4+\cdots,
\label{RG2}
\end{equation}
a form typical of a super-renormalizable field theory.%
\footnote{$b_1$ depends on the renormalization scheme:
It is not invariant under a perturbative change in scheme, $g^2=g'^2+bg'^4+\cdots$.
Hence we do not specify it further, until we adopt the Schr\"odinger functional scheme below.
Since $b_1$ comes from screening by physical particles, it should be positive in any sensible scheme.}
When $N_f$ is small, there is a region around $g^2=0$ where the running is dominated by the first term, which has the sign associated with asymptotic freedom in four dimensions.
As a consequence, the infrared running is towards strong coupling.
When the coupling becomes large enough, but before the second term becomes important, a condensate $\svev{\bar\psi\psi}$ forms.
This generates a mass for the fermions, whereupon at sufficiently small energy scales the fermions drop out of the theory, leaving free photons. The running of $e^2$ stops.
When $N_f$ is large, on the other hand, the one-loop term in \Eq{RG2} has to be considered.%
\footnote{In the 't~Hooft limit, $N_f\to\infty$ with $e^2N_f$ fixed, the one-loop term in \Eq{RG1} is the exact beta function.}
The beta function starts out negative at small $g^2$ but the one-loop term then makes it cross zero at $g^2=g_*^2\equiv(N_fb_1)^{-1}$.
This is an infrared-attractive fixed point.
Once the coupling runs there, a scale invariance sets in.
There are no massive particles and indeed, since Green functions behave as powers of distance, there are no particles at all.
There must be a value of $N_f$, which we have called $N_c$, that divides the two domains.
For $N_f$ just below this $N_c$, the one-loop term is significant: It make the beta function turn towards zero as $g^2$ grows.
This only happens, however, at a strong coupling, and the condensate is triggered before $g^2$ actually reaches the zero of the one-loop beta function.
Then the fermions develop a mass and decouple in the infrared, as described for the small-$N_f$ case.
These infrared scenarios are familiar from non-Abelian theories in four dimensions.
In an asymptotically free theory the one-loop term in the beta function is negative, while the two-loop term changes sign as the number of fermion flavors is increased~\cite{Caswell:1974gg,Banks:1981nn}.
The transition from confinement to conformality has been much studied, most recently by the methods of lattice gauge theory (for surveys see~\cite{Neil:2012cb,Giedt:2012it}).
The Schr\"odinger functional (SF) method gives a non-perturbative definition of the beta function that lends itself to lattice calculations~\cite{Luscher:1992an,Luscher:1993gh}.
Originally designed for QCD \cite{Sint:1995ch,Jansen:1998mx,DellaMorte:2004bc}, it has also been applied to gauge theories that are on the borderline between confining and conformal \cite{Appelquist:2007hu,Shamir:2008pb,Appelquist:2009ty,Hietanen:2009az,Bursa:2009we,Bursa:2010xn,Hayakawa:2010yn,Karavirta:2011zg,DeGrand:2010na,DeGrand:2011qd,DeGrand:2012yq,DeGrand:2012qa,DeGrand:2013uha}.
In this paper we present a study of QED3 with $N_f=2$ four-component fermions by this method.
We first calculate and plot results for the beta function at fixed lattice spacing.
These indicate that the beta function deviates from the one-loop form (\ref{RG2}) at strong coupling and does not cross zero.
This is confirmed when we extrapolate to the continuum by two different methods.
Unless this trend is reversed at yet stronger couplings, our results imply that the $N_f=2$ theory is a confining theory with a mass scale.
As mentioned above, more straightforward lattice methods have also been applied to QED3.
We note in particular the extensive studies of Refs.~\cite{Hands:2002dv,Hands:2004bh,Strouthos:2008kc}.
This work has pointed towards $N_c\approx2$.
Calculating mass spectra and the chiral condensate is then quite challenging in the $N_f=2$ theory, because of the strong finite-volume effects inherent in low dimensionality: If $N_c$ is nearby then the condensate (if any) could be very small and particle masses (if any) likewise.
Deciding whether these quantities are really nonzero requires simulation on very large lattices~\cite{Gusynin:2003ww}.
The strength of the SF method, on the other hand, comes from the idea behind the renormalization group.
There is no need to control a large range of length scales in a given calculation, because the method relates nearby scales to derive the beta function.
Thus finite size turns from a hindrance to a basis for calculation.
This paper is organized as follows.
In Sec.~\ref{sec:SF} we review the SF method by adapting it to QED3;
the language is entirely that of the continuum.
We describe the lattice theory in Sec.~\ref{sec:lattice}: a non-compact Abelian gauge field coupled to $N_f$=2 four-component fermions.
We simulate this theory to obtain the running coupling, with results presented in Sec.~\ref{sec:beta}.
Section~\ref{sec:conclusions} contains further discussion of our method, while the appendices contain some technical details.
\section{Defining the running coupling\label{sec:SF}}
The Schr\"odinger functional offers a definition of the running coupling
that is convenient for non-perturbative lattice calculations.
We apply the method to QED3 in the continuum; lattice modifications will be given below.
The action is
\begin{equation}
S=\int d^3x\,\left(\frac1{4e_0^2}F_{\mu\nu}F^{\mu\nu}
+\sum_{i=1}^{N_f}\bar\psi_i \Sl D\psi_i\right),
\label{cont_action}
\end{equation}
where $F_{\mu\nu}=\partial_\mu A_\nu - \partial_\nu A_\mu$ is the usual field strength and $D_\mu=\partial_\mu+iA_\mu$.
With this definition, $A$ has dimensions of mass and so does the coupling $e_0^2$.
Renormalizing at a scale $\mu$, we identify $e_0^2\equiv e^2(\mu)$ and
define a dimensionless coupling $g^2(\mu)=e^2(\mu)/\mu$.
We work in a cubic volume of size $L^3$ and impose a background field in a way
that $L$ is its only scale.
Then calculation of the quantum effective action yields a coupling that runs with $L$, which we denote by $g^2(L)$.
We impose the background field through boundary conditions on the spacelike components $A_x$ and $A_y$ at
$t=0$ and $t=L$,
\begin{eqnarray}
A_x &=& A_y \ = \ +\phi/L \,, \qquad t=0 \,,
\nonumber\\
A_x &=& A_y \ = \ -\phi/L \,, \qquad t=L \,.
\label{bc}
\end{eqnarray}
The background field in the bulk is found by minimizing
the classical action,
\begin{equation}
S_{\text{cl}}=\frac1{e_0^2}\int d^3x\,\frac14 F_{\mu\nu}F^{\mu\nu},
\label{classical_action}
\end{equation}
and it is easy to see that the solution is a linear function of $t$,
\begin{equation}
A_x(t)=A_y(t)=\frac\phi L\left(1-\frac{2t}L\right),
\label{Afield}
\end{equation}
corresponding to a constant electric field,
\begin{equation}
E_x=E_y=-\frac{2\phi}{L^2}.
\label{Efield}
\end{equation}
The classical action of this field is
\begin{equation}
S_{\text{cl}}=\frac{4\phi^2}{e_0^2L}\equiv \frac{\tilde{K}(\phi)}{e_0^2L}.
\label{defK}
\end{equation}
In our work we fixed the background field parameter to be $\phi=\pi/4$.
The effective action $\Gamma=-\log Z$ gives a definition of the
running coupling in the SF scheme.
We write
\begin{equation}
\Gamma \equiv \frac1{e^2(L)}\int d^3x\,\frac14 F_{\mu\nu}F^{\mu\nu},
\end{equation}
where $F_{\mu\nu}$ is the classical background field.
Using Eqs.~(\ref{classical_action}) and~(\ref{defK}) gives
\begin{equation}
\Gamma=\frac{\tilde{K}(\phi)}{e^2(L)L}=\frac{\tilde{K}(\phi)}{g^2(L)}\,,
\end{equation}
and thus a calculation of $\Gamma$, in a three-dimensional volume of size $L$, gives directly the running coupling $g^2(L)$ at the scale $L$.
In any field theory, one generally calculates not $\Gamma$ but its derivatives, which are given by Green functions.
Thus we differentiate with respect to the boundary parameter to obtain
\begin{equation}
\frac{\partial\Gamma}{\partial\phi}=\frac{K(\phi)}{g^2(L)}\,,
\label{GammaK}
\end{equation}
where $K = \partial\tilde{K}/\partial\phi$.
(Our choice $\phi=\pi/4$ gives $K=8\phi=2\pi$.)
We can check our numerical results by comparing to the first two terms in
a loop expansion for $\Gamma$,
\begin{equation}
\Gamma=S_{\text{cl}}+\Gamma^{(1)}+\cdots.
\end{equation}
The one-loop quantum correction is given by
\begin{equation}
\Gamma^{(1)}=-N_f{\rm tr}\,\log(\Sl D),
\end{equation}
and thus we define the one-loop quantity $c(\phi)$ via
\begin{equation}
\frac{\partial\Gamma^{(1)}}{\partial\phi}\equiv N_fK(\phi)c(\phi).
\label{NKc}
\end{equation}
[We have taken out a factor of $K(\phi)$ for convenience.]
Here $c(\phi)$ is a dimensionless function of the boundary conditions,
independent of the coupling and of the system size $L$.
We calculate it in Appendix \ref{app:oneloop}.
Inserting in \Eq{GammaK} we have the perturbative expression for the
running coupling,
\begin{equation}
\frac1{g^2(L)}=\frac1{g^2(\mu)\mu L}+N_fc+\cdots.
\label{RGsoln}
\end{equation}
Setting $q=1/L$, we rewrite the renormalization group equation (\ref{RG2}) as
\begin{equation}
\tilde\beta(1/g^2)\equiv\frac{d(1/g^2)}{d\log L}=-\frac1{g^2}+N_fb_1 +O(g^2).
\label{RG3}
\end{equation}
Upon differentiating \Eq{RGsoln}, we identify
\begin{equation}
b_1=c,
\end{equation}
the one-loop coefficient in the SF renormalization scheme.
\section{The lattice theory\label{sec:lattice}}
We use a non-compact formulation of the gauge field, wherein we define
the vector potential $A_{n\mu}$ on each link $(n,\mu)$ of the three-dimensional
lattice; we put a four-component
Dirac field $\psi_n$ on each site $n$.
(We suppress the flavor index $f=1,2$ throughout.)
The Euclidean action contains the usual quadratic term for the gauge field and
a smoothed Wilson--clover action for the fermions,
\begin{equation}
S=\frac{\beta}2\sum_{\scriptstyle n\atop\scriptstyle \mu<\nu}(\nabla\times A)_{n\mu\nu}^2
+\bar\psi D\psi
\label{action}
\end{equation}
All the fields in \Eq{action} have been made dimensionless via appropriate
powers of the lattice spacing $a$.
The lattice curl is
\begin{equation}
(\nabla\times A)_{n\mu\nu}=A_{n\mu}+A_{n+\hat\mu,\nu}-A_{n+\hat\nu,\mu}-A_{n\nu},
\label{curl}
\end{equation}
and the first summation in \Eq{action} counts each plaquette once.
The fermion term is
\begin{subequations}
\begin{eqnarray}
\bar\psi D\psi&=&\sum_n\bar\psi_n\psi_n\label{action-1}\\
&&+\kappa\sum_{n\mu}\left[\bar\psi_n(1+\gamma_\mu)V_{n\mu}\psi_{n+\hat\mu}\right.\nonumber\\
&&\quad\left.+\bar\psi_{n+\hat\mu}(1-\gamma_\mu)V_{n\mu}^\dag\psi_n\right]\label{action-2}\\
&&+\kappac_{\text{SW}}\sum_n\bar\psi_n\frac i4\sigma_{\mu\nu}F_{n\mu\nu}\psi_n.\label{action-3}
\end{eqnarray}
\label{actionF}
\end{subequations}
The Wilson hopping term (\ref{action-2}) contains a link connection $V_{n\mu}$.
This is constructed from the compact gauge variables
\begin{equation}
U_{n\mu}=e^{iA_{n\mu}}
\end{equation}
by a normalized hypercubic (nHYP) smearing process~ \cite{Hasenfratz:2001hp,Hasenfratz:2007rf}, where each $V_{n\mu}$ is a weighted average of
the $U$ variables on nearby links (see Appendix \ref{app:smear}).
The purpose of this smearing is the suppression of lattice artifacts.
It allows us to go to stronger couplings before encountering numerical instabilities~\cite{DeGrand:2010na}.
A further cancellation of lattice artifacts is offered by the clover term (\ref{action-3})~\cite{Sheikholeslami:1985ij}.
While the field strength $F_{n\mu\nu}$ could be defined via the simple curl
(\ref{curl}), we adopt a definition appropriate to the compact theory,
a sum over the four leaves of the ``clover'' surrounding the site $n$,
\begin{equation}
F_{n\mu\nu}=\frac14\left(F^{(1)}_{n\mu\nu}+F^{(1)}_{n-\hat\mu,\mu\nu}+F^{(1)}_{n-\hat\nu,\mu\nu}+F^{(1)}_{n-\hat\mu-\hat\nu,\mu\nu}\right),
\end{equation}
where each term $F^{(1)}$ is a compact curl,
\begin{equation}
F^{(1)}_{n\mu\nu}=\sin(\nabla\times A)_{n\mu\nu}.
\end{equation}
This clover structure is the same as in non-Abelian theories and enables easy
adaptation of existing code.
We set the coefficient to its tree-level value, $c_{\text{SW}}=1$, since we have found
this to be close to optimal when nHYP smearing is used~\cite{Shamir:2010cq}.
The boundary conditions (\ref{bc}) for the Schr\"odinger functional are now imposed on the gauge field on the time slices of the lattice at $t=0$ and $t=L$.
Moreover, there are no dynamical fermion fields on these boundaries.
The coupling in \Eq{action} is
\begin{equation}
\beta=\frac1{e_0^2a},
\end{equation}
where $e_0$ is the bare charge.
The hopping parameter is related to the bare electron mass $m_0$ by
\begin{equation}
\kappa=\frac1{6+2m_0a}.
\end{equation}
We study the massless theory by demanding that the measured, physical
fermion mass $m$ vanish.
We define this from the axial Ward identity,
\begin{equation}
\partial_\mu^- A_\mu^a=2m P^a,
\label{AWI}
\end{equation}
where $A^a_\mu=\frac12\bar\psi\gamma_\mu\gamma_5\tau^a\psi$ is the isovector axial current and $P^a=\frac12\bar\psi\gamma_5\tau^a\psi$ is the isovector
pseudoscalar density. ($\partial^-_\mu$ is the backward
lattice derivative.)
It is convenient in a SF calculation to define an gauge-invariant, pseudoscalar
wall source operator ${\cal O}^a$ near the boundary at $t=0$ (see Refs.~\cite{Shamir:2010cq,Luscher:1996sc} for details).
Then \Eq{AWI} can be used to relate two Green functions at zero spatial
momentum, viz.,
\begin{equation}
\partial_0^-\sum_{\bf x}\svev{A^a_0({\bf x},t){\cal O}^a}
=2m\sum_{\bf x}\svev{P^a({\bf x},t){\cal O}^a}.
\end{equation}
We evaluate the Green functions at $t=L/2$, whence the ratio gives $m$.
At each value of $\beta$, we tune the hopping parameter $\kappa$ to make $m$
vanish, which defines the critical curve $\kappa_c(\beta)$.
We list the values of $\beta$ used in this study, as well as the values of $\kappa_c(\beta)$, in Table~\ref{tab:Kc}.
As mentioned above, a Monte Carlo simulation of the lattice theory does not give the effective action directly.
Instead, one applies the Schr\"odinger functional method by using Green functions, which are derivatives of $\Gamma$.
The derivative on the left-hand side of Eq.~(\ref{GammaK}) is
\begin{equation}
\left.\frac{\partial \Gamma}{\partial\phi} \right|_{\phi=\pi/4}
=
\left.\svev{\frac{\partial S_{G}}{\partial\phi}
-{\rm tr}\, \left( \frac{1}{D^\dagger}\;
\frac{\partial (D^\dagger D)}{\partial\phi}\;
\frac{1}{D} \right)}\right|_{\phi=\pi/4},
\label{dphi}
\end{equation}
where $S_G$ is the pure gauge action, the first term in \Eq{action}.
Equation (\ref{dphi}) is a particular expectation value of the gauge fields and the Dirac operator $D$.
Differentiating with respect to $\phi$
is the same as differentiating
with respect to $A_{n\mu}$ on the boundary.
We also impose twisted spatial boundary conditions on the fermion fields~\cite{Sint:1995ch},
$\psi(x+L)=\exp(i\theta)\psi(x)$, with $\theta=\pi/5$ on both spatial axes.
We employed the hybrid Monte Carlo (HMC) algorithm~\cite{Duane:1985hz,Duane:1986iw,Gottlieb:1987mq} in our simulations. The molecular dynamics integration was accelerated with an
additional heavy pseudo-fermion field~\cite{Hasenbusch:2001ne}, multiple time
scales~\cite{Urbach:2005ji}, and a second-order Omelyan
integrator~\cite{Takaishi:2005tz}.
Since the non-compact formulation allows gauge fluctuations in which $A_{n\mu}$ can wander to infinity, we monitor the field and carry out gauge transformations, local and global, to keep the field within certain bounds.
\begin{table}
\begin{ruledtabular}
\begin{tabular}{lcrrrr}
& & \multicolumn{4}{c}{trajectories (thousands)}\\
\cline{3-6}
$\beta$ & $\kappa_{c}$ & $L=8a$ & $L=12a$ & $L=16a$ & $L=24a$\\
\hline
0.355 & 0.17440 & 150 & 150 & -- & --\\
0.39 & 0.17325 & 150 & 100 & -- & --\\
0.4 & 0.17282 & 100 & 150 & 150 & 122\\
0.458 & 0.17166 & 100 & 100 & -- & --\\
0.541 & 0.17055 & 100 & 100 & -- & --\\
0.6 & 0.16994 & 100 & 100 & 150 & 110\\
0.8 & 0.16898 & 100 & 100 & 100 & 100\\
1 & 0.16848 & 100 & 100 & 100 & 105\\
2 & 0.16757 & -- & -- & 50 & 50
\end{tabular}
\end{ruledtabular}
\caption{Summary of simulation runs for obtaining the running coupling $g^{2}$
at bare couplings $\left(\beta,\kappa_{c}\right)$ for lattice sizes
$L$ used in this work. We used a trajectory length of unity for most
of the simulations. The exceptions are the strong coupling runs ($\beta\le0.4$)
and the $\beta=0.6$ run on the biggest lattice, in which we used a
trajectory length of $1/2$.
\label{tab:Kc} }
\end{table}
\begin{table}
\begin{ruledtabular}
\begin{tabular}{lcccc}
& \multicolumn{4}{c}{$1/g^{2}$}\\
\cline{2-5}
$\beta$ & $L=8a$ & $L=12a$ & $L=16a$ & $L=24a$\\
\hline
0.355 & 0.0753(20) & 0.0515(35) & -- & --\\
0.39 & 0.0751(18) & 0.0608(42) & -- & --\\
0.4 & 0.0753(24) & 0.0591(34) & 0.0558(41) & 0.0422(75)\\
0.458 & 0.0824(22) & 0.0644(41) & -- & --\\
0.541 & 0.0992(16) & 0.0771(21) & -- & --\\
0.6 & 0.1110(16) & 0.0833(22) & 0.0686(21) & 0.0485(63)\\
0.8 & 0.1374(17) & 0.1018(22) & 0.0854(26) & 0.0635(32)\\
1 & 0.1643(16) & 0.1258(21) & 0.0978(25) & 0.0837(32)\\
2 & -- & -- & 0.1732(35) & 0.1277(43)
\end{tabular}
\end{ruledtabular}
\caption{Schr\"odinger functional running coupling calculated at bare couplings $\beta$ on lattices of size $L$. The first two lines are affected by lattice artifacts that make them unusable in calculating the RDBF.
\label{tab:schrodinger-func}}
\end{table}
\section{The running coupling and the beta function \label{sec:beta}}
We calculated the running coupling from Eqs.~(\ref{GammaK}) and~(\ref{dphi}) for lattice sizes
$L/a=8,$ 12, 16, and 24 for the bare couplings listed in Table~\ref{tab:Kc}.
The results are shown in Table \ref{tab:schrodinger-func}.
We use different subsets of the data for two different analysis methods.
Our goal is the beta function $\tilde\beta$, defined in \Eq{RG3}, or, equivalently, its representation as the rescaled discrete beta function (RDBF) \cite{DeGrand:2011qd} for scaling by a fixed factor $s$,
\begin{equation}
R(u,s)\equiv\frac{u(sL)-u(L)}{\log s},
\label{RDBF}
\end{equation}
where the argument is
\begin{equation}
u=u(L)\equiv\frac{1}{g^{2}\left(L\right)}.
\end{equation}
It is clear that at a fixed point, where $\tilde\beta=0$, the RDBF will be zero as well.
\subsection{Discrete beta function---two-lattice matching}
We calculate the RDBF directly by comparing pairs of lattice sizes $L$ and $L'=sL$ at fixed bare coupling $\beta$.
[Two lattices with the same $(\beta,\kappa)$ have the same lattice spacing.]
Lattices of size $L=8a$ and~$12a$ give a scale factor $s=3/2$, as do lattices of size $L=16a$ and $24a$.
We plot the RDBF for all such pairs of lattices in Fig.~\ref{fig:RDBF}.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth,clip]{RDBF.eps}
\end{center}
\caption{The rescaled discrete beta function calculated from each pair of couplings for the
lattices of size $L/a=8\rightarrow12$ and $16\rightarrow24$.
\label{fig:RDBF}}
\end{figure}
For comparison we plot the one-loop formula, derived from using \Eq{RGsoln} in the definition (\ref{RDBF}),
\begin{equation}
R^{(1)}(u,s)=\frac{1-1/s}{\log s}\left(-u+N_fc\right).
\end{equation}
We note the general trend that as $u$ decreases the data deviate downwards from the one-loop curve---away from the axis---and avoid its zero.
Figure~\ref{fig:RDBF} is a first look only, since the dependence on lattice spacing has not yet been studied.
The quantity plotted in Fig.~\ref{fig:RDBF} is really $R(u,s;a)$, where the added argument is the lattice spacing.
To extrapolate to $a=0$, we seek data for $R$ at fixed $u$---which means fixed $L$---but at different $a$, and thus different lattice size $L/a$.
We show this procedure in Fig.~\ref{fig:two-lattice}.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth,clip]{a3dcouplings_L.eps}
\end{center}
\caption{Extrapolating the RDBF to the continuum.
Plotted is a subset of the results for the running coupling that are listed in Table~\ref{tab:schrodinger-func}.
Pairs of data points at the same $\beta$ give the RDBF for a single lattice spacing.
The horizontal lines link data points at the same coupling $u=1/g^2$ but at different lattice spacings.
These pairs give the RDBF at the two lattice spacings, to be compared and extrapolated in Fig.~\ref{fig:RDBF2}.
\label{fig:two-lattice}}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth,clip]{RDBF_extrap.eps}
\end{center}
\caption{Extrapolation of the RDBF to the continuum, at the two couplings for which we have the needed data.
\label{fig:RDBF2}}
\end{figure}
The horizontal lines link data points at fixed $u$, but measured on different lattice sizes $L/a=8$ and~16.%
\footnote{The points for $L=8a$ were calculated at $\beta=0.458$ and~$0.541$---lines~4
and~5 of Table~\ref{tab:schrodinger-func}.
These $\beta$ values were chosen for this purpose, i.e., to match $u$ to the $L=16a$ values at $\beta=0.8$ and~1.}
Thus the RDBFs calculated from these points are $R(u,s;a=L/8)$ and $R(u,s;a=L/16)$, respectively.
We have such pairs at fixed $u$ at $u\simeq0.84$ and $u\simeq0.98$.
We replot them in Fig.~\ref{fig:RDBF2}, together with the extrapolations according to
\begin{equation}
R(u,s;a)=R(u,s;a=0)+C\frac aL.
\end{equation}
It proved impossible to extend the two-lattice matching method to stronger coupling because of lattice artifacts.
We attempted to extend the $L=8a$ data to stronger couplings than those shown in Fig.~\ref{fig:two-lattice} (see Table~\ref{tab:schrodinger-func}).
The measured coupling $1/g^2$ levels off, rather than continuing downward with smaller $\beta$.
This prevents matching to the $L=16a$ data.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth,clip]{a3dcouplings.eps}
\end{center}
\caption{Running couplings $1/g^2$ vs $\log(L/a)$, from Table~\ref{tab:schrodinger-func}.
Each set corresponds to a fixed bare coupling $\beta$.
Top to bottom: $\beta=1$, 0.8, 0.6,~0.4.
The solid lines are the linear fits (\ref{l1fit}) for
each bare coupling.
The horizontal, dashed lines are values of $1/g^2$ chosen for extrapolating the slopes to $a/L=0$.
\label{fig:sfB1}}
\end{figure}
\subsection{Slope analysis}
In order to perform an analysis at stronger couplings than the two-lattice method allows, we begin by plotting data sets for $1/g^2$ in Fig.~\ref{fig:sfB1}.
At four values of $\beta$, we have results for $1/g^2$ on four lattice volumes, $L=8a$, $12a$, $16a$, and~$24a$.
Since fixing $\beta$ fixes the lattice spacing $a$, these sets represent scaling in $L$ at fixed $a$.
As $L$ is increased, $1/g^2$ decreases in accord with a negative beta function.
The figure shows the results of a straight-line fit to each data set,
\begin{equation}
u(L)=\frac{1}{g^{2}(L)}=c_{0}+c_{1}\log L/a.
\label{l1fit}
\end{equation}
At the two weakest bare couplings, $\beta=1.0$ and~0.8, the fit is poor because of the obvious curvature of the dependence on $\log (L/a)$.
This means that the decrease in $1/g^2$ is not described by a constant beta function $\tilde\beta$.
Indeed, the one-loop continuum formula predicts that the decrease in $1/g^2$ slows as one moves leftward in Fig.~\ref{fig:RDBF}.
At $\beta=0.6$ and~0.4, on the other hand, a straight line fits the points well, with perhaps a small deviation for $L=24a$ where the error bar is large anyway.
Since $g^2$ is {\em not\/} constant as $L$ changes, this motivates the hypothesis that the beta function $\tilde\beta(u)$ has leveled off in the range of $u$ covered by these data.
Thus we arrive at estimates $\tilde\beta(u)\simeq c_1$.
The different slopes of the two data sets show that the estimates vary with the lattice spacing.
Choosing a value of $u=1/g^2$ in Fig.~\ref{fig:sfB1} gives a horizontal line that intersects the fit lines for the two data sets.
The two intercepts give two values of $a/L$, one for each fit.
We again extrapolate these to the continuum linearly,
\begin{equation}
\tilde\beta(u;a)=\tilde\beta(u;a=0)+C\frac aL,
\end{equation}
as shown in Table~\ref{tab:slopes} and Fig.~\ref{fig:tbeta_ex}.
The final result is plotted in Fig.~\ref{fig:tbeta}.
The three $u$-values chosen span a short interval.
We do not go to yet stronger coupling in this analysis because lowering the chosen value of $u$ in Fig.~\ref{fig:sfB1} will give intercepts that are too far to the right; here the data deviate from the straight lines, as they must, and the straight lines will not give accurate values of $\tilde\beta$.
[A complementary argument is the following. The quality of the linear fits in Fig.~\ref{fig:sfB1}
is good for both $\beta=0.4$ and 0.6.
Therefore, when we extrapolate to the continuum limit
it is legitimate to use the two slopes as estimates for the beta functions
for any value $0.0422 \le u \le 0.0753$,
which is the interval covered by the $\beta=0.4$ data.
(It is contained in the interval covered by the $\beta=0.6$ data.)
As we move down in $u$, however, the intercepts with the linear fits
get closer to each other, and, moreover, the uncertainty in the $L/a$ value of each intercept grows. As a result, the uncertainty in
the continuum extrapolation grows rapidly, a trend that is clear in Figs.~\ref{fig:tbeta_ex} and~\ref{fig:tbeta}.
Thus no further information will emerge from pushing to stronger coupling.]
Figure~\ref{fig:tbeta} shows that our numerical results deviate considerably from the one-loop curve.
They show no sign of crossing the axis;
taking Figs.~\ref{fig:RDBF2} and~\ref{fig:tbeta} together, we have a beta function that approaches the axis with the one-loop curve but then curves away from it in the strong coupling region.
\begin{table}
\begin{ruledtabular}
\begin{tabular}{llll}
$1/g^2$ & $L/a(\beta=0.4)$ & $L/a(\beta=0.6)$ & $\tilde\beta(a/L\to0)$\\
\hline
0.0753(24) & \ 8 & 14.10(64) & $-0.103(13)$ \\
0.0686(21) & \ 9.73(87) & 16 & $-0.11(2)$ \\
0.06 & 12.85(96) & 18.04(65) & $-0.14(3)$
\end{tabular}
\end{ruledtabular}
\caption{Extrapolation to $a/L=0$ in slope analysis.
$1/g^2$ is the chosen value of the running coupling for the horizontal line in Fig.~\ref{fig:sfB1}.
The next two columns are the values of $L/a$ for the intercepts of the horizontal line with the fit lines for $\beta=0.4$ and~0.6.
The slopes for the two lines, which give $\tilde\beta(u;a)$, are respectively -0.031(6) and -0.062(4).
The final column gives the extrapolation from these two values to $a/L=0$ (Figs.~\ref{fig:tbeta_ex} and~\ref{fig:tbeta}).
\label{tab:slopes}}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth,clip]{slope_extrap.eps}
\end{center}
\caption{Extrapolating the beta function to the continuum, as described in Table~\ref{tab:slopes}.
The $y$-values of the upper and lower sets of data are the slopes of the fits shown in Fig.~\ref{fig:sfB1} for $\beta=0.4$ and $0.6$.
The $x$-values are the inverses of the intercepts of the horizontal lines in that figure with the fit lines.
The data are extrapolated to $a/L=0$ as shown (solid points, displaced horizontally from the axis for clarity).
\label{fig:tbeta_ex}}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth,clip]{slope_method.eps}
\end{center}
\caption{Extrapolation of the beta function to the continuum (Table~\ref{tab:slopes} and Fig.~\ref{fig:tbeta_ex}), plotted against the coupling.
The points labeled $\beta=0.4$ and~0.6 are the $y$-values of the data plotted in Fig.~\ref{fig:tbeta_ex}.
Assuming the slopes in Fig.~\ref{fig:sfB1} to be constant in the coupling range shown gives a set of extrapolations to $a/L=0$, depending on the chosen value of $u$, as shown.
The extrapolations are not statistically independent of each other.
\label{fig:tbeta}}
\end{figure}
\section{Discussion \label{sec:conclusions}}
QED3 is qualitatively different from the four-dimensional non-Abelian theories studied with the SF method~\cite{Appelquist:2007hu,Shamir:2008pb,Appelquist:2009ty,Hietanen:2009az,Bursa:2009we,Bursa:2010xn,Hayakawa:2010yn,Karavirta:2011zg,DeGrand:2010na,DeGrand:2011qd,DeGrand:2012yq,DeGrand:2012qa,DeGrand:2013uha}.
This is evident already in the weak-coupling behavior of the beta function.
In four dimensions, the beta function $\tilde\beta(u)$ for the inverse coupling is constant in one loop; in three, it is linear in $u$, as seen in \Eq{RG3} and in the figures.
Moreover, in the borderline-conformal theories there is a partial cancellation between the one-loop and two-loop terms that makes the beta function small compared to that of QCD.
These differences are reflected in the analysis technique we adopt here, which is different from that in our earlier papers
\cite{DeGrand:2011qd,DeGrand:2012yq,DeGrand:2012qa,DeGrand:2013uha}
and more closely resembles the applications of the method to QCD~\cite{DellaMorte:2004bc}.
Our technique of linear fits at fixed bare coupling $\beta$ [\Eq{l1fit} and Fig.~\ref{fig:sfB1}] was designed for a slowly running theory, where the slopes are small and the coupling changes little from the smallest lattice to the largest.
It works best when the beta function changes slowly as well.
It is plain that neither property holds in Fig.~\ref{fig:sfB1}:
the slopes are large and the couplings change rapidly.
For the two largest values of $\beta$, where the beta function is largest, the rate of change of the coupling decreases markedly as one follows the data to large volumes, so that the non-constancy of the beta function is evident.
At the smaller $\beta$'s the slopes hold more nearly constant, validating the fits and allowing us to make the hypothesis that the beta function has levelled off.
We have used two different methods for extrapolation to the continuum, each with its own limitations.
In weak coupling, we used the two-lattice method to extrapolate the rescaled discrete beta function.
This works as long as one can match couplings between two different lattice sizes.
As we saw, it fails at stronger couplings because of lattice artifacts.
At the stronger bare couplings we use the fact that, as in QCD, the data in Fig.~\ref{fig:sfB1} cover overlapping ranges in the running coupling.
The overlap between $\beta=0.6$ and~0.4 allows for a straightforward extrapolation of the slope---the beta function---to $a/L=0$.
It is evident from Fig.~\ref{fig:sfB1} that the slopes decrease as $\beta$ decreases, as is to be expected from the form of the one-loop beta function.
Thus the horizontal lines in Fig.~\ref{fig:sfB1} will associate the larger lattices (in terms of $L/a$) with larger slopes.
It is then inevitable that extrapolating the slopes in Fig.~\ref{fig:tbeta_ex} pushes the beta function away from the axis.
Is the result, then, due more to our method than to the data?
The examples offered by the four-dimensional non-Abelian theories show that this is not the case~\cite{DeGrand:2011qd,DeGrand:2012yq,DeGrand:2012qa,DeGrand:2013uha}.
There, as one approaches a possible zero of the beta function, the slopes become so small that the fixed-$\beta$ data sets do not overlap in coupling.
This prevents an extrapolation of slopes to the continuum limit in those theories.
An alternative method, based only the linear fits at fixed $\beta$, leads to a beta function that stays near (or even crosses) zero~\cite{DeGrand:2012yq}.
\begin{acknowledgments}
We thank N.~Itzhaki for urging us to work on QED3.
B.S. thanks the UCLA Physics Department for its hospitality when this work was begun.
This work was supported in part by the Israel Science Foundation under Grants~423/09
and~1362/08 and by the European Research Council under Grant~203247.
Our lattice simulation program had its origins in the publicly available package of the MILC collaboration
\cite{MILC} with nHYP smearing adapted from code by A.~Hasenfratz, R.~Hoffmann, and S.~Schaefer~\cite{Hasenfratz:2001tw}.
\end{acknowledgments}
|
\section{}
\tableofcontents
\section{Introduction}
The most studied example of AdS/CFT is $\mathcal{N}=4$ super Yang-Mills theory with $U(N)$ gauge group, and its dual, type IIB string theory on AdS$_{5}\times S^{5}$. Testing the AdS/CFT correspondence in non-planar limits of the gauge theory and its dual string theory is a very interesting problem. In this regard, integrability would be a powerful tool. Therefore, it is interesting to investigate if the integrability, that was found in the planar limit \cite{Beisert}, is also present in non-planar limits. Furthermore, testing AdS/CFT in less supersymmetric settings is also an interesting problem. In particular we have the $1/4$-BPS sector of the theory in mind. The study of restricted Schur polynomials has yielded progress in both these directions. A restricted Schur is a local operator in the gauge theory which can be built from a variety of fields. For instance, we can use the complex scalar Higgs fields, the fermion fields as well as the gauge fields to build these operators. We consider restricted Schurs built using two types of complex scalar fields, $Z$ and $Y$ say. In this case, the definition is
\begin{equation}
\label{eq:RSUNDef}
\chi_{R(r,s)\alpha\beta}(Z,Y) = \frac{1}{n!m!}\sum\limits_{\sigma \in S_{n+m}} \mathrm{Tr}\big( P_{R\rightarrow (r,s)\alpha\beta} \Gamma_{R}(\sigma) \big) \mathrm{Tr}\big( \sigma Z^{\otimes n}\otimes Y^{\otimes m} \big)
\end{equation}
\\
In (\ref{eq:RSUNDef}), $R$ is a Young diagram with $n+m$ boxes, corresponding to an irreducible representation (irrep) of $S_{n+m}$ and $(r,s)$ are a pair of Young diagrams with $n$ and $m$ boxes respectively corresponding to an irrep of $S_{n}\times S_{m}$. $P_{R\rightarrow (r,s)\alpha\beta}$ is a projector in the labels $R$ and $(r,s)$ and an intertwiner in the labels $\alpha$ and $\beta$. These labels are the multiplicity labels with which irrep $(r,s)$ is subduced from $R$ when restricting from $S_{n+m}$ to $S_{n}\times S_{m}$. Further details of (\ref{eq:RSUNDef}) can be found in \cite{GGWSA1}, \cite{GGO} for example. There are many good reasons to study the operators defined in (\ref{eq:RSUNDef}) as we now explain.
The counting of states given by the partition function of the $1/4$-BPS sector of the free $U(N)$ theory was shown to match the counting of restricted Schurs that can be defined \cite{Collins}. Essentially, the partition function was expressed in terms of Littlewood Richardson coefficients which count the number of restricted Schurs for a given set of Young diagram labels, $R,(r,s)$. This result was generalised to an arbitrary product of $U(N)$ gauge groups in \cite{Quiver}.
We are interested in operators whose bare dimension grows parametrically with $N$. For these operators, the large $N$ limit of correlation functions is not captured by summing only planar diagrams \cite{GGO}. The usual $1/N^{2}$ factor suppressing non-planar diagrams is over-powered by combinatorial factors resulting from evaluating all possible Wick contractions. Indeed, when the number of matrix fields in the operator scales with $N$, we are forced to sum an infinite number of non-planar diagrams in the large $N$ limit. A solution to this problem was given in \cite{Jevicki} for the $1/2$-BPS case. In this important and influential work, it was observed that there is a one-to-one correspondence between the space of $1/2$-BPS representations and Young diagrams, i.e., Schur polynomials. Using representation theory of the symmetric and unitary group, the two-point function of Schur polynomials was computed \emph{exactly} in the free theory limit. \cite{EMC} then achieved the remarkable result of computing the two-point function of restricted Schurs (\ref{eq:RSUNDef}) exactly in the free theory. The operators diagonalised the two-point function and a very simple formula for the final result was given. Thus, the restricted Schurs provide an exactly orthogonal basis for the $1/4$-BPS sector of the free $U(N)$ gauge theory.
Apart from the interesting properties of (\ref{eq:RSUNDef}) as gauge theory objects, restricted Schurs also have an AdS/CFT dual. When the scaling dimension of operators grows with $N$, $\chi_{R(r,s)\alpha\beta}(Z,Y)$, with $R$ having long columns or rows ($O(N)$ boxes), has been argued to be dual to a system of excited giant gravitons \cite{Invasion}, \cite{GGO}. Concretely, such a system is realised as a system of giant gravitons with strings attached \cite{GGWSA1}. Thus, restricted Schurs are a useful tool for studying non-perturbative physics of both the gauge theory and its dual string theory.
In a similar approach, \cite{Kimura:2007wy} wrote down an orthogonal basis of gauge invariant operators using $Z$ and $Z^{\dagger}$ using Brauer algebras and their permutation sub-algebras in the free theory limit. \cite{Ram} then also studied $1/4$ and $1/8$-BPS operators at zero coupling, $g^{2}_{YM} = 0$. Also by making heavy use of representation theory, the multi-matrix multi-trace operators constructed here were argued to form a basis and to diagonalise their two-point function. When the operators in \cite{Ram} consist of two matrix fields $Z$ and $Y$ say, they were shown to be identical to the basis of restricted Schur polynomials \cite{Collins}. Progress has also been made in the counting and constructing of $1/4$ and $1/8$-BPS operators at weak coupling \cite{Weakcoup}.
Tremendous progress has already been achieved in the study of the anomalous dimension spectrum of restricted Schur polynomials; see \cite{GGO}, \cite{NPI}, \cite{Spring}, \cite{twoloop}, \cite{HigherLoopAnDim} for examples. Anomalous dimensions of the restricted Schurs are dual to excitation energies of the excited giant graviton system. One of the main results of these works is that the spectrum of anomalous dimensions reduces to the set of normal mode frequencies of a system of decoupled harmonic oscillators. The fact that we get decoupled oscillators is evidence of integrability in the non-planar sectors studied in these works.
Is there a similar story for the theory with an $SO(N)$ gauge group? In this work, we take the first steps toward answering this question. $\mathcal{N} = 4$ super Yang-Mills theory with an $SO(N)$ gauge group is dual to type IIB string theory with AdS$_{5}\times \mathcal{R}$P$^{5}$ geometry. At the non-planar level, the spectral problems of $U(N)$ and $SO(N)$ gauge theories are distinct \cite{SpecProbSON}. Thus, there is a good chance that a study of anomalous dimensions of large dimension operators will reveal new aspects of D-brane physics. The $1/2$-BPS case for free $SO(N)$ gauge theory was studied in \cite{SON1}, \cite{SON2} and \cite{Diaz} . The results of \cite{SON1} and \cite{SON2} include defining Schur polynomials in the square of the eigenvalues and computing their two-point functions exactly in the free theory. Here too, the Schur polynomial was shown to diagonalise the two-point function, a property that was argued to be independent of the gauge group \cite{Diaz}.
We show that restricted Schur polynomials provide an exactly orthogonal basis for the $1/4$-BPS sector of free super Yang-Mills theory with $SO(N)$ gauge group. In this work, we take $N$ to be even. First we show that the counting of states obtained from the partition function in this sector matches the number of restricted Schurs that can be defined. This is achieved by expressing the partition function in terms of the Littlewood Richardson coefficients. The counting comes out differently from the $1/4$-BPS $U(N)$ case as we will see. $SO(N)$ has two invariant tensors. They are $\delta_{ij}$ and $\epsilon_{i_{1}i_{2}\cdots i_{N}}$. We focus on operators constructed using $\delta_{ij}$. We construct restricted Schur polynomials and manage to compute the two-point function exactly in the free theory. The result is expressed in a relatively simple formula. In the next section, we present some background theory needed for our computations.
\section{Projectors}
In this section, we briefly discuss some of the projectors and states that appear frequently in our calculations. All projectors we discuss here project from the carrier space of irrep $R$ of $S_{2q}$, $q=n+m$, onto the carrier space of an irrep of some subgroup.
\subsection{From $R$ onto $(r,s)$}
Firstly we discuss the projectors $P_{R\rightarrow (r,s)\alpha\beta}$ constructed in \cite{GGO}. They project from $R$ onto irrep $(r,s)$ of the subgroup $S_{2n}\times S_{2m}$. In the multiplicity labels, however, $P_{R\rightarrow (r,s)\alpha\beta}$ is an intertwiner \cite{RepTh}. We write these operators as
\begin{equation}
\label{eq:1}
P_{R\rightarrow (r,s)\alpha\beta} = \sum\limits^{d_{r}\times d_{s}}_{l=1} \ket{R(r,s)\alpha;l}\!\!\bra{R(r,s)\beta;l}
\end{equation}
\\
where $\ket{R(r,s)\alpha;l}$ is a state in the carrier space of the $\alpha$-th copy of irrep $(r,s)$ and $d_{r}\times d_{s}$ is the dimension of $(r,s)$. In this basis, the matrix representation of $\gamma \in S_{2n}\times S_{2m}$ in irrep $R$ is block diagonal with the $S_{2n}\times S_{2m}$ irreps as the diagonal blocks. For example
\begin{equation}
\label{eq:2}
\Gamma_{R}(\gamma) = \left(\begin{array}{cccc}\Gamma_{(r',s')}(\gamma) & 0 & 0 & 0 \\0 & \Gamma_{(r'',s'')}(\gamma) & 0 & 0 \\0 & 0 & \Gamma_{(r,s)1}(\gamma) & 0 \\0 & 0 & 0 & \Gamma_{(r,s)2}(\gamma)\end{array}\right)
\end{equation}
\\
where we have shown two copies of irrep $(r,s)$. There are four projectors that can be defined for $(r,s)$. Acting on $\Gamma_{R}(\gamma)$ each give the following
\begin{eqnarray}
P_{R\rightarrow(r,s)11}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & \Gamma_{(r,s)1}(\gamma) & 0 \\0 & 0 & 0 & 0\end{array}\right)\\
P_{R\rightarrow(r,s)22}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & \Gamma_{(r,s)2}(\gamma)\end{array}\right)\\
P_{R\rightarrow(r,s)12}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & \Gamma_{(r,s)2}(\gamma) \\0 & 0 & 0 & 0\end{array}\right)\\
P_{R\rightarrow(r,s)21}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & \Gamma_{(r,s)1}(\gamma) & 0\end{array}\right)
\end{eqnarray}
\\
\subsection{From $R$ onto $[S]$}
Let $R$ have an even number of boxes in each row. Restricting $S_{2q}$ to the subgroup $S_{q}[S_{2}]$, one can find a basis in which $\Gamma_{R}(\xi)$, $\xi\in S_{q}[S_{2}]$, is block diagonal with the $S_{q}[S_{2}]$ irreps appearing on the diagonal. For $R$ having even rows, there exists a 1-dimensional irrep of $S_{q}[S_{2}]$ which we denote $[S]$. In this irrep, all matrices are represented by 1. Let $\ket{[S]}$ be the state spanning the carrier space of $[S]$. Then we have \cite{Ivanov}
\begin{equation}
\label{eq:3}
\Gamma_{R}(\xi)\ket{[S]} = \ket{[S]}, \hspace{20pt} \xi \in S_{q}[S_{2}].
\end{equation}
\\
We can also define a projector to go from $R$ to $[S]$.
\begin{equation}
\label{eq:4}
P_{[S]} = \frac{1}{q!2^{q}}\sum\limits_{\xi\in S_{q}[S_{2}]}\Gamma_{R}(\xi)
\end{equation}
\\
The state $\ket{[S]}$ may be calculated as the eigenvector of $P_{[S]}$ with eigenvalue 1. Lastly, $[S]$ is subduced from $R$ with multiplicity 1 \cite{SON1}.
\subsection{From $(r,s)\beta$ to $([A],[A])\beta$}
Recall that $(r,s)\beta$ is the $\beta$-th copy of the irrep of $S_{2n}\times S_{2m}$. If $r$ has an even number of boxes in each column, then, upon restricting to the $S_{n}[S_{2}]$ subgroup another type of 1 dimensional irrep may be subduced \cite{Ivanov}. Call this irrep $[A]$. In this irrep,
\begin{equation}
\label{eq:5}
\Gamma_{r}(\eta) = \mathrm{sgn}(\eta), \hspace{20pt} \eta \in S_{n}[S_{2}].
\end{equation}
\\
Denote by $\ket{[A]}$ the state spanning the 1-dimensional carrier space of $[A]$. Irrep $[A]$ is also subduced from $r$ with no multiplicity \cite{SON1}. If $r$ and $s$ both have even columns, then each may subduce the irrep $[A]$ of $S_{n}[S_{2}]$ and $S_{m}[S_{2}]$ respectively. Thus, $(r,s)\beta$ subduces the irrep $([A],[A])\beta$ of $S_{n}[S_{2}]\times S_{m}[S_{2}]$. Denote by $\ket{[A],[A]\beta}$ the state spanning this 1-dimensional irrep. Define
\begin{equation}
\label{eq:6}
P_{R\rightarrow ([A],[A])\beta} = P_{R\rightarrow(r,s)\beta\beta}P_{[A,A]}, \hspace{20pt} P_{[A,A]} = \frac{1}{n!m!2^{q}}\sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]}\mathrm{sgn}(\eta)\Gamma_{R}(\eta)
\end{equation}
\\
The state $\ket{[A],[A]\beta}$ may be defined to be the eigenvector of $P_{R\rightarrow ([A],[A])\beta}$ with eigenvalue 1. This state has the following properties
\begin{enumerate}
\item \begin{equation}
\label{eq:611}
P_{R\rightarrow (r,s)\alpha\beta} \ket{[A],[A]\beta} = \ket{[A],[A]\alpha}
\end{equation}
\item
\begin{equation}
\Gamma_{R}(\eta) \ket{[A],[A]\beta} = \mathrm{sgn}(\eta)\ket{[A],[A]\beta},\hspace{20pt} \eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]
\end{equation}
\item The quantity $\langle[A],[A]\beta | R(r,s)\beta;l\rangle $ does not depend on the multiplicity index $\beta$. This will be important when we construct the restricted Schurs.
\end{enumerate}
\section{Counting}
In this section we express the partition function for the $1/4$-BPS sector of free $SO(2n)$ gauge theory in terms of the Littlewood Richardson coefficients depending on three Young diagram labels. Two Young diagrams will have an even number of boxes in each \emph{column}, and the third one (induced from the previous two) will have an even number of boxes in each \emph{row}. This is first demonstrated for the $SO(4)$ case for simplicity. It is then easy to extend the argument to the general $SO(2n)$ case, which follows thereafter. The Littlewood Richardson number $g(\lambda,\mu,\xi)$ counts the number of times a Young diagram $\xi$ is induced from two smaller Young diagrams $\mu$ and $\lambda$, say. Thus, $g(\lambda,\mu,\xi)$ counts the number of restricted Schurs that may be defined for labels $\{\lambda,\mu,\xi\}$.
The partition function for the $1/4$-BPS sector of free $SO(2n)$ gauge theory is \cite{SON1}, \cite{Dolan1}
\begin{eqnarray}
\label{eq:SO2nPF}
G(t_{1},t_{2}) &=& \frac{1}{2^{n-1}n!}\int_{T_{n}} \prod\limits^{n}_{j=1} \frac{dx_{j}}{2\pi ix_{j} }\Delta(x+x^{-1})^{2}\times \\
&& \prod\limits^{2}_{k=1}\prod\limits_{1\leqslant i<j\leqslant n}\frac{1}{(1-t_{k})^{n}}\frac{1}{(1-t_{k}x_{i}x_{j})(1-t_{k}x_{i}x^{-1}_{j})(1-t_{k}x^{-1}_{i}x_{j})(1-t_{k}x^{-1}_{i}x^{-1}_{j})} \nonumber
\end{eqnarray}
\\
Let's study the simple $SO(4)$ case first and generalise to $SO(2n)$ thereafter. For $SO(4)$, this becomes
\begin{eqnarray}
\label{eq:1steqnforG}
G(t_{1},t_{2}) &=& \frac{1}{4}\int \frac{dx_{1}}{2\pi ix_{1}}\int \frac{dx_{2}}{2\pi ix_{2}}\Delta(x+x^{-1})^{2}\times \\
&& \prod\limits^{2}_{k=1}\frac{1}{(1-t_{k})^{2}}\frac{1}{(1-t_{k}x_{1}x_{2})(1-t_{k}x_{1}x^{-1}_{2})(1-t_{k}x^{-1}_{1}x_{2})(1-t_{k}x^{-1}_{1}x^{-1}_{2})}\nonumber
\end{eqnarray}
\\
The factor in the second line may be written as
\begin{eqnarray}
&&\frac{1}{(1-t_{k}x_{1}x^{-1}_{1})(1-t_{k}x_{2}x^{-1}_{2})(1-t_{k}x_{1}x_{2})(1-t_{k}x_{1}x^{-1}_{2})(1-t_{k}x^{-1}_{1}x_{2})(1-t_{k}x^{-1}_{1}x^{-1}_{2})} \nonumber \\
\label{eq:rewritingprod}
&&\hspace{200pt} =\; \prod\limits_{1\leqslant i< j \leqslant 4}\frac{1}{1-t_{k}y_{i}y_{j}}
\end{eqnarray}
\\
with $y_{1}=x_{1}, y_{2} = x^{-1}_{1}, y_{3} = x_{2}, y_{4} = x^{-1}_{2}$. Now we may expand this product in terms of Schur polynomials using the formula \cite{Ellitic}
\begin{equation}
\label{eq:modcaulittd}
\prod\limits_{1\leqslant i< j \leqslant L}\frac{1}{1-y_{i}y_{j}} = \sum\limits_{\mu}s_{\mu^{2}}(y_{1},..., y_{L})
\end{equation}
\\
where $\mu^{2}$ is defined as the partition with parts
\begin{equation}
\label{eq:7}
(\mu^{2})_{i} = \mu_{\lceil i/2 \rceil}.
\end{equation}
\\
This means the following. Take, for example, $i = \{1,2,3,4\}$. Then, $i/2 = \{ 1/2, 1, 3/2, 2 \}$. The ceiling symbol then tells us to take: $\lceil i/2 \rceil = \{ 1,1,2,2 \}$. So, if $\mu^{2}\,_{i}$ denotes the $i$th part of $\mu^{2}$, then
\begin{equation}
\label{eq:8}
\mu^{2} = (\mu^{2}\,_{1},\mu^{2}\,_{2},\mu^{2}\,_{3},\mu^{2}\,_{4}) = (\mu_{1},\mu_{1},\mu_{2},\mu_{2})
\end{equation}
\\
Thus, in general for a partition $\mu^{2}$ with $n$ parts,
\begin{equation}
\label{eq:9}
\mu^{2} = (\mu^{2}\,_{1},\mu^{2}\,_{2},\mu^{2}\,_{3},\mu^{2}\,_{4},\cdots, \mu^{2}\,_{n-1},\mu^{2}\,_{n}) = (\mu_{1},\mu_{1},\mu_{2},\mu_{2},\cdots,\mu_{n/2},\mu_{n/2})
\end{equation}
\\
Partitions $\mu^{2}$ then correspond to Young diagrams with an even number of boxes in each column. Including the prefactor $t_{k}$ in (\ref{eq:modcaulittd}), we get
\begin{equation}
\label{eq:911}
\prod\limits_{1\leqslant i< j \leqslant L}\frac{1}{1-t_{k}y_{i}y_{j}} = \sum\limits_{\mu}s_{\mu^{2}}(\sqrt{t_{k}}y_{1},..., \sqrt{t_{k}}y_{L}) = \sum\limits_{\mu} (t_{k})^{|\mu^{2}|/2} s_{\mu^{2}}(y_{1},...,y_{L})
\end{equation}
\\
where we used the fact that Schur polynomial $s_{\lambda}(y_{1},y_{2},\cdots, y_{L})$ may be written as
\begin{equation}
\label{eq:10}
s_{\lambda}(ty_{1},ty_{2},\cdots,ty_{L}) = t^{|\lambda|}s_{\lambda}(y_{1},y_{2},\cdots, y_{L}).
\end{equation}
\\
In the above formulas, $|\,|$ stands for the weight of the partition. The partition function (\ref{eq:1steqnforG}) becomes
\begin{eqnarray}
\label{eq:11}
G(t_{1},t_{2}) &=& \frac{1}{4}\int \frac{dx_{1}}{2\pi ix_{1}}\int \frac{dx_{2}}{2\pi ix_{2}}\Delta(x+x^{-1})^{2}\times \\
&&\sum\limits_{\lambda}\sum\limits_{\mu}(t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2}s_{\lambda^{2}}(x_{1},x^{-1}_{1},x_{2},x^{-1}_{2})s_{\mu^{2}}(x_{1},x^{-1}_{1},x_{2},x^{-1}_{2})\nonumber
\end{eqnarray}
\\
Now, we use the product rule involving Schur polynomials. This means that we can write the product of two Schur polynomials with partitions $\lambda^{2}$ and $\mu^{2}$ as the sum over a single Schur with partition $\xi$ times by the Littlewood Richardson coefficient $g(\lambda^{2},\mu^{2},\xi)$ \cite{Dolan2}. The Littlewood Richardson coefficient is how many times $\xi$ subduces $(\lambda^{2},\mu^{2})$. Thus,
\begin{eqnarray}
\label{eq:12}
G(t_{1},t_{2}) &=& \sum\limits_{\xi}\sum\limits_{\lambda}\sum\limits_{\mu} g(\lambda^{2},\mu^{2},\xi) (t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2} \times\\
&& \frac{1}{4}\int \frac{dx_{1}}{2\pi ix_{1}}\int \frac{dx_{2}}{2\pi ix_{2}}\Delta(x+x^{-1})^{2}s_{\xi}(x_{1},x^{-1}_{1},x_{2},x^{-1}_{2}) \nonumber
\end{eqnarray}
\\
Next, we generalise this to $SO(2n)$. In the same way as in equation (\ref{eq:rewritingprod})
\begin{equation}
\label{eq:13}
\frac{1}{(1-t_{k})^{n}}\prod\limits_{1\leqslant i <j \leqslant n}\frac{1}{(1-t_{k}x_{i}x_{j})(1-t_{k}x_{i}x^{-1}_{j})(1-t_{k}x^{-1}_{i}x_{j})(1-t_{k}x^{-1}_{i}x^{-1}_{j})} = \prod\limits_{1\leqslant i< j \leqslant N}\frac{1}{1-t_{k}y_{i}y_{j}}
\end{equation}
\\
with $y_{i} = x_{\lceil i/2 \rceil}$ and $y_{2i} = x^{-1}_{i}$ for $i = 1,2,\cdots, n$. Using equation (\ref{eq:modcaulittd}) for each factor,
\begin{equation}
\label{eq:14}
\prod\limits^{2}_{k=1} \prod\limits_{1\leqslant i< j \leqslant N}\frac{1}{1-t_{k}y_{i}y_{j}} = \sum\limits_{\lambda}\sum\limits_{\mu}(t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2}s_{\lambda^{2}}(y_{1},..,y_{N})s_{\mu^{2}}(y_{1},..,y_{N})
\end{equation}
\\
After using the product rule again, the $SO(2n)$ partition function in (\ref{eq:SO2nPF}) becomes
\begin{eqnarray}
\label{eq:InteformofPF}
G(t_{1},t_{2}) &=& \sum\limits_{\xi}\sum\limits_{\lambda}\sum\limits_{\mu} g(\lambda^{2},\mu^{2},\xi) (t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2} \times\\
&& \frac{1}{2^{n-1}n!}\int_{T_{n}} \prod\limits^{n}_{j=1} \frac{dx_{j}}{2\pi ix_{j}}\Delta(x+x^{-1})^{2}s_{\xi}(x_{1},x^{-1}_{1},...,x_{n},x^{-1}_{n})\nonumber
\end{eqnarray}
\\
Recognising the Haar (or $G$-invariant) measure for $SO(2n)$, the integral in (\ref{eq:InteformofPF}) may be written as
\begin{equation}
\label{eq:15}
I_{n} = \int_{SO(2n)}[dO] \,s_{\xi}(\mathrm{x})
\end{equation}
\\
This integral is equal to 1 for two cases. The first is if $\xi$ has an even number of boxes in each \emph{row}. The second is if $\xi$ has a Young diagram with an even number of boxes in each row stuck onto a single column of $2n$ boxes\footnote{Recall that there are only two invariant tensors for $SO(2n)$. They are $\delta_{ij}$ and $\epsilon_{i_{1}i_{2}\cdots i_{2n}}$. This case is relevent for operators that are built using $\epsilon_{i_{1}i_{2}\cdots i_{2n}}$.}. Here are examples of each for $SO(4)$
\begin{equation}
\label{eq:16}
\xi = \begin{Young} &&&\cr&\cr&\cr \end{Young}, \hspace{20pt} \mathrm{or} \hspace{20pt} \xi = \begin{Young} &&\cr&&\cr\cr\cr \end{Young}
\end{equation}
\\
The integral vanishes in all other cases. See \cite{Ellitic} \cite{Vanishing} and \cite{SymmFunc}. We have checked these results for many examples for $SO(4)$ and $SO(6)$. With this result, the $SO(2n)$ partition function becomes
\begin{equation}
\label{eq:FinalPF}
G(t_{1},t_{2}) = \sum\limits_{\xi}\sum\limits_{\lambda}\sum\limits_{\mu} g(\lambda^{2},\mu^{2},\xi) (t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2}
\end{equation}
\\
where $\xi$ is a Young diagram with even rows and $\lambda^{2}$ and $\mu^{2}$ are Young diagrams with even columns. This result is truly different from the $U(N)$ $1/4$-BPS case, studied in \cite{Collins}. For $U(N)$ the counting went according to the square of the Littlewood-Richardson number. Here, the counting goes according to the Littlewood Richardson number itself. In other words, for a given $S_{2n}\times S_{2m}$ irrep, $(r,s)$, each with even columns, and given $S_{2q}$ irrep $R$ with even rows, the number of restricted Schurs one can define is exactly $g(r,s,R)$. We present some counting examples in appendix 1.
\section{Defining the restricted Schurs}
We now discuss the problem of defining the restricted Schurs for the $1/4$-BPS free $SO(N)$ gauge theory. We focus on operators that can be constructed using $\delta$ rather than $\epsilon_{i_{1}i_{2}\cdots i_{2n}}$. We construct these operators out of two complex matrix fields $Z$ and $Y$. The two-point function of these fields is given by \cite{SON1}
\begin{equation}
\label{eq:17}
\langle Z^{ij}\overline{Z}^{kl} \rangle = \langle Z^{ij}Z_{kl} \rangle = \delta^{i}_{k}\delta^{j}_{l} - \delta^{i}_{l}\delta^{j}_{k}
\end{equation}
\\
and similarly for $Y$. We define our $SO(N)$ restricted Schurs to be
\begin{equation}
\label{eq:RSDef}
O_{R(r,s)\alpha}(Z,Y) = \frac{1}{(2n)!(2m)!}\sum\limits_{\sigma \in S_{2q}} \mathrm{Tr}\big( \mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)C^{4\nu}_{I}\sigma^{I}_{J} (Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}
\end{equation}
\\
where
\begin{equation}
\label{eq:ODef}
\mathcal{O}_{R(r,s)\alpha} = \ket{[S]}\!\!\bra{[A],[A]\beta}P_{R\rightarrow (r,s)\beta\alpha}
\end{equation}
\\
In (\ref{eq:RSDef}), the tensor $(Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}$ is defined as
\begin{equation}
\label{eq:18}
(Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J} = Z^{j_{1}j_{2}}\cdots Z^{j_{2n-1}j_{2n}}Y^{j_{2n+1}j_{2n+2}}\cdots Y^{j_{2q-1}j_{2q}}
\end{equation}
\\
$C^{4\nu}_{I}$ is the contractor defined in \cite{SON2}
\begin{equation}
\label{eq:19}
C^{4\nu}_{I} = \delta_{k_{1}k_{2}}\cdots \delta_{k_{2q-1}k_{2q}}(\sigma_{4\nu})^{K}_{I}
\end{equation}
\\
$C^{4\nu}_{I}$ is responsible for contracting the free indices in such a way to make the operator gauge invariant. For $\sigma_{4\nu}$, we take
\begin{equation}
\label{eq:20}
\sigma_{4\nu} = (1,2,3,4)(5,6,7,8)\cdots (2q-3,2q-2,2q-1,2q)
\end{equation}
\\
This permutation contracts indices $1\&4$, $2\&3$ $\cdots$, $2q-3\&2q$ and $2q-2\&2q-1$, i.e.,
\begin{equation}
\label{eq:21}
C^{4\nu}_{I}(\sigma)^{I}_{J} = \sigma^{i_{1}i_{2}i_{2}i_{1}\cdots i_{q-1}i_{q}i_{q}i_{q-1}}_{j_{1}j_{2}j_{3}j_{4}\cdots j_{2q}}.
\end{equation}
\\
$R$ is an irrep of $S_{2q}$ and $(r,s)$ is an irrep of the subgroup $S_{2n}\times S_{2m}$. Indices $\alpha$ and $\beta$ are multiplicity labels for the irrep $(r,s)$. State $\ket{[A],[A]\beta}$ is the state spanning the one-dimensional carrier space $([A],[A])\beta$ of $S_{n}[S_{2}]\times S_{m}[S_{2}]$ as defined in section 2. Concretely, it is the eigenvector of the operator $P_{R\rightarrow ([A],[A])\beta}$ with eigenvalue 1. State $\ket{[S]}$ is the state spanning the 1-dimensional carrier space $[S]$ of $S_{q}[S_{2}]$ as defined in section 2. The $S_{q}[S_{2}]$ here is chosen to stabilise the set $(1,4),(2,3),\cdots, (2q-3,2q),(2q-2,2q-1)$. Concretely, $\ket{[S]}$ is the eigenvector of $P_{[S]}$.
The coefficients in (\ref{eq:RSDef}) satisfy the following property
\begin{equation}
\label{eq:22}
\mathrm{Tr}\big( \mathcal{O}_{R(r,s)\alpha} \Gamma_{R}(\sigma) \big)\mathrm{sgn}(\eta) = \mathrm{Tr}\big( \mathcal{O}_{R(r,s)\alpha} \Gamma_{R}(\eta\sigma\xi) \big), \hspace{20pt} \xi \in S_{q}[S_{2}], \eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]
\end{equation}
\\
Lastly, the operators in (\ref{eq:RSDef}) depend only on a single multiplicity label and thus, match the counting found in section 3.
\section{Two-point function}
In this section, we evaluate the two-point function of the operators defined in section 4. Recall the definition for our restricted Schurs,
\begin{eqnarray}
O_{R(r,s)\alpha} &=& \frac{1}{(2n)!(2m)!}\sum\limits_{\sigma\in S_{2q}}\mathrm{Tr}(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R})C^{4\nu}_{I}\sigma^{I}_{J}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}\\
\overline{O}_{T(t,u)\beta} &=& \frac{1}{(2n)!(2m)!}\sum\limits_{\sigma\in S_{2q}}\mathrm{Tr}(\mathcal{O}_{T(t,u)\beta}\Gamma_{R})C^{K}_{4\nu}(\overline{\sigma})^{L}_{K}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})_{L}
\end{eqnarray}
\\
where
\begin{equation}
\label{eq:23}
(\overline{\sigma})^{J}_{I} = \delta^{j_{\sigma(1)}}_{i_{1}}\delta^{j_{\sigma(2)}}_{i_{2}}\cdots \delta^{j_{\sigma(2q)}}_{i_{2q}} = \delta^{j_{1}}_{i_{\sigma^{-1}(1)}}\delta^{j_{2}}_{i_{\sigma^{-1}(2)}}\cdots \delta^{j_{2q}}_{i_{\sigma^{-1}(2q)}} = (\sigma^{-1})^{J}_{I}.
\end{equation}
\\
The first step in the calculation is evaluating the correlator
\begin{equation}
\label{eq:CorrZY}
\langle (Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})_{L} \rangle
\end{equation}
\\
In \cite{SON1} the correlator $(\langle Z^{\otimes 2n})^{J}(Z^{\otimes 2n})_{L} \rangle$ was found to be a sum over permutations belonging to the wreath product $S_{n}[S_{2}]$, where each term in the sum was weighted by the $\mathrm{sgn}$ of the permutation. The correlator in (\ref{eq:CorrZY}) generalises to a sum over the subgroup $S_{n}[S_{2}]\times S_{m}[S_{2}]$
\begin{equation}
\label{CorrZYanswer}
\langle (Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})_{L} = \sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]} \mathrm{sgn}(\eta) \eta^{J}_{L}
\end{equation}
\\
This formula may be easily checked for the example of $n=m=2$.
\begin{eqnarray}
&&\langle Z^{i_{1}i_{2}}Z^{i_{3}i_{4}}Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Z}^{j_{1}j_{2}}\bar{Z}^{j_{3}j_{4}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle = \nonumber \\
&&\langle Z^{i_{1}i_{2}}\bar{Z}^{j_{1}j_{2}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{3}j_{3}} \rangle \Big[ \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{5}j_{6}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{7}j_{8}} \rangle + \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{7}j_{8}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}} \rangle \Big] +\nonumber\\
&& \langle Z^{i_{1}i_{2}}\bar{Z}^{j_{3}j_{4}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{1}j_{2}} \rangle \Big[ \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{5}j_{6}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{7}j_{8}} \rangle + \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{7}j_{8}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}} \rangle \Big]\nonumber\\
&=& \langle Z^{i_{1}i_{2}}\bar{Z}^{j_{1}j_{2}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{3}j_{3}} \rangle \langle Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle + \langle Z^{i_{1}i_{2}}\bar{Z}^{j_{3}j_{4}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{1}j_{2}} \rangle \langle Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle \nonumber\\
&=& \langle Z^{i_{1}i_{2}}Z^{i_{3}i_{4}}\bar{Z}^{j_{1}j_{2}}\bar{Z}^{j_{3}j_{4}} \rangle \langle Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle \nonumber\\
&=& \bigg[\sum\limits_{\rho\in S_{2}[S_{2}]}\mathrm{Sgn}(\rho)\rho\bigg] \bigg[\sum\limits_{\sigma\in S_{2}[S_{2}]}\mathrm{Sgn}(\sigma)\sigma\bigg]\nonumber\\
&=& \sum\limits_{\psi \in S_{2}[S_{2}]\times S_{2}[S_{2}]}\mathrm{Sgn}(\psi)\psi^{I}_{J}
\end{eqnarray}
\\
where $\psi^{I}_{J} = \rho \sigma$ and $\mathrm{Sgn}(\psi)=\mathrm{Sgn}(\rho)\mathrm{Sgn}(\sigma)$. Now compute the two-point function
\begin{eqnarray}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle &=& \frac{1}{\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\sigma,\rho\in S_{2q}}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)\mathrm{Tr}\big(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho)\big)C^{4\nu}_{I}C^{L}_{4\nu}\sigma^{I}_{J}(\rho^{-1})^{K}_{L} \nonumber \\
&& \times \langle (ZY)^{J}(ZY)_{K} \rangle \nonumber \\
&=& \frac{1}{\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\sigma,\rho\in S_{2q}}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)\mathrm{Tr}\big(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho)\big)C^{4\nu}_{I}C^{L}_{4\nu}\sigma^{I}_{J}(\rho^{-1})^{K}_{L} \nonumber \\
&& \sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]}\mathrm{sgn}(\eta)\eta^{J}_{K} \nonumber \\
&=& \frac{1}{\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\sigma,\rho\in S_{2q}}\sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)\mathrm{Tr}\big(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho)\big)\mathrm{sgn}(\eta) \nonumber \\
\label{eq:TPF1}
&& \hspace{50pt} \times \;C^{4\nu}_{I}C^{L}_{4\nu}(\rho^{-1}\eta\sigma)^{I}_{L}
\end{eqnarray}
\\
Relabelling the sum over $\sigma = \eta^{-1}\psi$ where $\psi \in S_{2q}$, we obtain
\\
\begin{equation}
\label{eq:TPF2}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \frac{n!m!2^{q}}{\big( (2n)!(2m)! \big)^{2}} \sum\limits_{\psi,\rho\in S_{2q}}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\psi)\big)\mathrm{Tr}(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho))C^{4\nu}_{I}C^{L}_{4\nu}(\rho^{-1}\psi)^{I}_{L}
\end{equation}
\\
We also used the fact that $\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]$, which is a subgroup of $S_{2n}\times S_{2m}$ and that
\begin{equation}
\label{eq:24}
\bra{[A],[A]\gamma}\Gamma_{R}(\eta^{-1}) = \bra{[A],[A]\gamma}\mathrm{sgn}(\eta^{-1})
\end{equation}
\\
which then canceled with $\mathrm{sgn}(\eta)$ in (\ref{eq:TPF1}). Next, relabel the sum over $\psi$ by letting $ \psi = \rho\tau$, with $\tau \in S_{2q}$. After using the orthogonality relation
\begin{equation}
\label{eq:25}
\sum\limits_{\rho \in S_{2q}} \Gamma_{R}(\rho)_{ij}\Gamma_{T}(\rho)_{kl} = \frac{(2q)!}{d_{R}}\delta_{RT}\delta_{ik}\delta_{jl}
\end{equation}
\\
and using
\begin{equation}
\label{eq:26}
\mathcal{O}_{R(r,s)\alpha}\mathcal{O}^{T}_{T(t,u)\beta} = \ket{[S]}\!\!\bra{[A],[A]\gamma}P_{R\rightarrow(r,s)\gamma\zeta} \ket{[A],[A]\zeta}\!\!\bra{[S]}\delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta},
\end{equation}
\\
where $T$ in the superscript is the transpose of $\mathcal{O}$, equation (\ref{eq:TPF2}) becomes
\begin{equation}
\label{eq:TPF3}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\tau\in S_{2q}}\mathrm{Tr}\big(\hat{P}_{[S]}\Gamma_{R}(\tau)\big)C^{4\nu}_{I}C^{L}_{4\nu}(\tau)^{I}_{L}
\end{equation}
\\
where we also used the fact that
\begin{equation}
\label{eq:27}
\bra{[A],[A]\gamma}P_{R\rightarrow(r,s)\gamma\zeta}\ket{[A],[A],\zeta} = 1
\end{equation}
\\
to simplify the two-point function. In (\ref{eq:TPF3}), the $S_{q}[S_{2}]$ in $\hat{P}_{[S]}$ is the stabiliser for $(1,4)(2,3)\cdots$. For reasons that will become clear shortly, we want to change the embedding of the $S_{q}[S_{2}]$. Let $\rho$ be a permutation such that
\begin{equation}
\label{eq:28}
C^{4\nu}_{I}(\rho)^{I}_{K} = C^{4\mu}_{K}.
\end{equation}
\\
where $\sigma_{4\mu} = (1,2)(3,4) \cdots (2q-1,2q)$. Defining
\begin{equation}
\label{eq:29}
\tau = \rho^{-1}\psi\rho
\end{equation}
\\
equation (\ref{eq:TPF3}) becomes
\begin{equation}
\label{eq:TPF4}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi\in S_{2q}}\mathrm{Tr}\big(\hat{P}_{[S]}\Gamma_{R}(\rho^{-1}\psi\rho)\big)C^{4\mu}_{K}C^{L}_{4\mu}(\psi)^{K}_{L}\\
\end{equation}
\\
With this relabelling, we are now contracting indices $1\&2$, $3\&4$ $\cdots (2q-1\&2q)$,
\begin{equation}
\label{eq:30}
C^{4\mu}_{K}C^{L}_{4\mu}(\psi)^{K}_{L} = (\psi)^{k_{1}k_{1}k_{2}k_{2}\cdots k_{q}k_{q}}_{l_{1}l_{1}l_{2}l_{2}\cdots l_{q}l_{q}}
\end{equation}
\\
Inside the trace, $\Gamma_{R}(\rho)\hat{P}_{[S]}\Gamma_{R}(\rho^{-1}) $ is now a sum over $S_{q}[S_{2}]$ which stabilises the set $(1,2),(3,4)\cdots (2q-1,2q)$. We now have the following for the two-point function
\begin{equation}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi\in S_{2q}}\mathrm{Tr}\big(P_{[S]}\Gamma_{R}(\psi)\big)C^{4\mu}_{K}C^{L}_{4\mu}(\psi)^{K}_{L}\\
\end{equation}
\\
Now split the sum over $S_{2q}$ up into a sum over the Brauer algrebra $\mathcal{B}_{q}$ and a sum over $S_{q}[S_{2}]$ as was done in \cite{SON1}. $\mathcal{B}_{q}$ is isomorphic to the coset $S_{2q}/S_{q}[S_{2}]$.
\begin{equation}
\label{eq:TPF5}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta}\frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi_{1}\in \mathcal{B}_{q}}\sum\limits_{\psi_{2}\in S_{q}[S_{2}]}\mathrm{Tr}\big(P_{[S]}\Gamma_{R}(\psi_{1}\psi_{2})\big)C^{4\mu}_{K}C^{L}_{4\mu}(\psi_{1}\psi_{2})^{K}_{L}
\end{equation}
\\
Now $\psi_{2}$ is an element of the stabiliser of $\sigma_{4\mu}$, which means that
\begin{equation}
\label{eq:31}
C^{4\mu}_{K}(\psi_{2})^{K}_{I} = C^{4\mu}_{I}
\end{equation}
\\
Then
\begin{equation}
\label{eq:32}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta}\frac{(2q)!q!n!m!2^{2q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi_{1}\in \mathcal{B}_{q}}\mathrm{Tr}\big(\Gamma_{R}(\psi_{1})P_{[S]}\big)C^{4\mu}_{I}C^{L}_{4\mu}(\psi_{1})^{I}_{L}
\end{equation}
\\
Since $\mathcal{B}_{q}$ is the set of all coset representatives, and therefore not unique, we again choose the following elements for $\mathcal{B}_{q}$ \cite{SON1}\footnote{Indeed, this is why we changed the embedding in the first place - so we could use the permutations in (\ref{eq:CosetReps}) as the set of coset representatives.}.
\begin{equation}
\label{eq:CosetReps}
\prod^{q-1}_{j=0}\prod^{2j+1}_{i=1}(i,2j+1)
\end{equation}
\\
It then follows that
\begin{equation}
\label{eq:TPFformula}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!q!n!m!2^{2q}}{d_{R}((2n)!(2m)!)^{2}}\bra{[S]}\big(\prod^{q-1}_{j=0}(N+J_{2j+1})\ket{[S]}
\end{equation}
\\
where we wrote $P_{[S]} = \ket{[S]}\!\!\bra{[S]}$. $J_{2j+1}$ is the Jucys-Murphy element in irrep $R$. $J_{2j+1}$ acting on a state in $R$ returns the content of the box labeled $2j+1$. Thus, $N + j_{2j+1}$ gives the weight of that box. Keep in mind that the $\ket{[S]}$ in (\ref{eq:TPFformula}) is symmetric in boxes $(1,2)$ and $(3,4) \cdots$, whereas the $\ket{[S]}$ in the definition of the restricted Schurs is symmetric in boxes $(1,4)$ and $(2,3)$ and so on.
The state $\ket{[S]}$ in (\ref{eq:TPFformula}) only consists of states in $R$ that has boxes $1\&2$ next to each other, and $3\&4$ next to each other $\cdots$ and boxes $2q-1\&2q$ next to each other. As examples, we have
\begin{equation}
\label{eq:33}
\begin{Young} 1&2&3&4\cr5&6\cr7&8\cr \end{Young} \hspace{20pt} and \hspace{20pt}\begin{Young} 1&2&5&6\cr3&4\cr7&8\cr \end{Young}
\end{equation}
\\
contributing to $\ket{[S]}$ for this particular $R$. In the end, it is only boxes in the odd columns whose weights contribute to the product in (\ref{eq:TPFformula}). By odd columns, we mean the weights of the all the boxes in the 1st column and the weights of all the boxes in the 3rd column and then the 5th column and so on. For example we take the weights of the starred boxes
\begin{equation}
\label{eq:34}
\begin{Young} *&\cr*&\cr \end{Young},\;\;\;\begin{Young} *&&*&\cr*&&*&\cr \end{Young}, \;\;\;\begin{Young} *&\cr*&\cr*&\cr*&\cr \end{Young}, \;\;\; \begin{Young} *&&*&\cr*&\cr*&\cr \end{Young}
\end{equation}
\\
We then arrive at the following formula for the two-point function
\begin{equation}
\label{eq:Finalformula}
\langle O_{R(r,s)\alpha} \overline{O}_{T(t,u)\beta}\rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!q!n!m!2^{2q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\prod\limits_{i\;\in\; odd\;columns\;in\;R} c_{i}
\end{equation}
\\
We test our formula for a variety of $S_{4}$ and $S_{8}$ examples in appendix 2. Unfortunately, the first multiplicity in the $S_{2n}\times S_{2m}$ irreps we found occurred for the case of
\begin{equation}
R = \begin{Young} &&&&&\cr&&&\cr&&&\cr&\cr&\cr&\cr & \cr \end{Young} \hspace{20pt} \mathrm{subducing\;2\;copies\;of} \hspace{20pt} (r,s) = ( \begin{Young} &&&\cr&&&\cr&\cr&\cr\cr \cr \end{Young}, \begin{Young} &&\cr&&\cr\cr \cr\end{Young})
\end{equation}
\\
which is beyond our computational capability.
\section{Discussion}
In this work, we have constructed an exactly orthogonal basis for the $1/4$-BPS sector of free $\mathcal{N}=4$ super Yang-Mills theory with $SO(2n)$ gauge group. We have shown that the counting of states in this sector exactly matches the number of restricted Schurs the can be defined for $(r,s)$ having even columns and $R$ having even rows. Furthermore, the counting went according to $g(r,s,R)$. This means that there are less operators than for the $U(N)$ case for which the counting went according to $g(r,s,R)^{2}$. We have also presented a variety of simple examples in support of our findings. We then constructed a basis of operators which matched the counting. The basis we constructed also has the nice symmetry property of being invariant under sending
\begin{equation}
\label{eq:35}
\sigma \rightarrow \eta \sigma \xi \hspace{20pt} \eta \in S_{n}[S_{2}]\times S_{m}[S_{2}] \; \mathrm{and}\; \xi \in S_{q}[S_{2}].
\end{equation}
\\
Finally we achieved an analytic formula for the two-point function in the form of (\ref{eq:Finalformula}). The restricted Schurs (\ref{eq:RSDef}) are orthogonal and its two-point function is relatively simple to evaluate. We take the product of the weights of the boxes in the odd columns. We also computed a few simple, but non-trivial, examples to check this formula. A next natural step is to study the spectrum of anomalous dimensions of these operators. Using AdS/CFT, we can then study physics of the dual brane objects. We hope to make progress in this direction soon.\\
\emph{Acknowledgements:} I would like to thank Pablo Diaz, Nirina Tahiridimbisoa for many useful discussions. In particular, I'd like to thank Prof. Robert de Mello Koch for all the input and encouragement, and without whom this work would certainly not have been possible.
\section{Appendix 1. Counting examples}
\subsubsection{1 $Z$ and 1 $Y$ for $SO(4)$}
First consider the case of having only 1 $Z$ and 1 $Y$ for $SO(4)$. The partition function gives a total of 2 operators. We now try to match this number with the number of restricted Schurs that can be defined. There is only 1 possible diagram for $r$,
\begin{equation}
\label{eq:36}
\begin{Young} \cr\cr \end{Young}
\end{equation}
\\
and one possible Young diagram for $s$,
\begin{equation}
\label{eq:37}
\begin{Young} \cr\cr \end{Young}
\end{equation}
\\
Multiplying $r$ and $s$ together, we find
\begin{equation}
\label{eq:38}
\begin{Young} \cr\cr \end{Young}\otimes \begin{Young} 1\cr2\cr \end{Young} =
\begin{Young}& 1\cr&2\cr \end{Young} \;\otimes \; \begin{Young} \cr\cr1\cr2\cr \end{Young}
\end{equation}
\\
Using $\delta_{ij}$ and $\epsilon_{ijkl}$, we can construct only are two operators:
\begin{eqnarray}
\label{eq:39}
\mathcal{O}_{1} &=& \mathrm{Tr}(ZY)
\label{eq:40}\\
\mathcal{O}_{2} &=& \epsilon_{ijkl}Z^{ij}Y^{kl}
\end{eqnarray}
\\
Thus, we have the following correspondence
\begin{eqnarray}
\mathrm{Tr}(ZY) &\leftrightarrow& \begin{Young}& \cr&\cr \end{Young}, (\begin{Young}\cr\cr\end{Young},\begin{Young}\cr\cr\end{Young})\\
\epsilon_{ijkl}Z^{ij}Y^{kl} &\leftrightarrow& \begin{Young} \cr\cr\cr\cr \end{Young}, (\begin{Young}\cr\cr\end{Young},\begin{Young}\cr\cr\end{Young})
\end{eqnarray}
\\
where the last operator is known as the Pfaffian \cite{SON1}
\subsubsection{2 $Z$'s and 2 $Y$'s for $SO(4)$}
For 2 $Z$'s and 2 $Y$'s for $SO(4)$, the partition function gives a total of 7 operators. We thus expect 7 different restricted Schur labels. For $r$ we can have
\begin{equation}
\label{eq:41}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
For $s$ we can have
\begin{equation}
\label{eq:42}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying $r$ and $s$ together and only taking the results with even rows, we get
\begin{eqnarray}
\begin{Young} &\cr&\cr \end{Young} \otimes \begin{Young} 1&1\cr2&2\cr \end{Young} &\mathrm{gives}&
\begin{Young} &&1&1\cr&&2&2\cr \end{Young}\;,\; \begin{Young} &&1&1\cr&\cr2&2\cr \end{Young} \;,\; \begin{Young} &&1\cr&&2\cr1\cr2\cr \end{Young}\;
\;,\;\begin{Young} &\cr&\cr 1&1 \cr2 &2\cr \end{Young}
\end{eqnarray}
\\
Next
\begin{eqnarray}
\begin{Young} &\cr&\cr \end{Young} \otimes \begin{Young} 1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr3\cr4\cr \end{Young}
\end{eqnarray}
\\
Then
\begin{eqnarray}
\begin{Young} \cr\cr\cr\cr \end{Young} \otimes \begin{Young} 1&1\cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr \end{Young}
\end{eqnarray}
\\
Finally
\begin{eqnarray}
\begin{Young} \cr\cr\cr\cr \end{Young} \otimes \begin{Young} 1\cr2\cr3\cr4\cr \end{Young}&\mathrm{gives}& \begin{Young} &1\cr&2\cr&3 \cr &4\cr \end{Young}
\end{eqnarray}
\\
Counting all the Young diagrams we get 7 with the correct labelling. We also check that we find 7 operators using $\delta_{ij}$ and $\epsilon_{ijkl}$. They are listed below
\begin{eqnarray}
\mathcal{O}_{1}&=& \mathrm{Tr}(Z^{2}Y^{2}) \nonumber \\
\mathcal{O}_{2}&=& \mathrm{Tr}(ZYZY) \nonumber \\
\mathcal{O}_{3}&=& \mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2})\nonumber \\
\mathcal{O}_{4}&=& \mathrm{Tr}(ZY)^{2} \nonumber \\
\mathcal{O}_{5} &=& \epsilon_{ijkl}Z^{ij}Z^{kl}\mathrm{Tr}(Y^{2}) \nonumber \\
\mathcal{O}_{6} &=& \epsilon_{ijkl}Z^{ij}Y^{kl}\mathrm{Tr}(ZY) \nonumber \\
\mathcal{O}_{7} &=& \epsilon_{ijkl}Y^{ij}Y^{kl}\mathrm{Tr}(Z^{2})
\end{eqnarray}
\\
\subsubsection{3 $Z$'s and 2 $Y$'s for $SO(6)$}
Here we do not expect multitrace operators. This is because there is an odd number of fields; we will always end up with a $\mathrm{Tr}(Z)$ or $\mathrm{Tr}(Y)$ which vanishes. We can, however, obtain operators built using the $\epsilon_{ijklmn}$. The partition function gives a total of 4 operators. We thus expect 4 restricted Schurs and each of the should have the Pfaffian as a factor. For $r$, the possible Young diagrams are
\begin{equation}
\label{eq:43}
\begin{Young} &&\cr&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
and for $s$, the possible Young diagrams are
\begin{equation}
\label{eq:44}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying each $r$ with each $s$, and taking only the diagrams which contribute, we find
\begin{eqnarray}
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& 0 \\
&& \nonumber \\
\begin{Young} &\cr&\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr\cr\cr1\cr2\cr \end{Young} \\
&& \nonumber \\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr\cr\cr \end{Young}\\
&&\nonumber \\
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&\cr&&\cr1\cr 2\cr3\cr4\cr \end{Young}\\
&& \nonumber \\
\begin{Young} &\cr&\cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr\cr\cr3\cr4\cr \end{Young}
\end{eqnarray}
\\
\begin{eqnarray}
&& \nonumber \\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}\otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0
\end{eqnarray}
\\
We indeed obtain 4 restricted Schurs as expected.
\subsubsection{4 $Z$'s and $2$Y's for SO(8)}
The partition function gives a total of 13 operators. The possible diagrams for $r$ are
\begin{equation}
\label{eq:45}
\begin{Young} &&&\cr&&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr&\cr&\cr \end{Young} , \hspace{20pt} \begin{Young} &&\cr&&\cr\cr \cr \end{Young} , \hspace{20pt} \begin{Young} &\cr&\cr\cr \cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
and for $s$
\begin{equation}
\label{eq:46}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying each $r$ with each $s$ and taking only the diagrams which contribute to the counting, we find
\begin{eqnarray}
\begin{Young} &&&\cr&&&\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&&&1&1\cr&&&&2&2\cr \end{Young} \;,\; \begin{Young} &&&&1&1\cr&&&\cr2&2\cr \end{Young}\;,\;\begin{Young} &&&\cr&&&\cr 1&1\cr2&2\cr\end{Young} \nonumber \\
\begin{Young} &&&\cr&&&\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0 \nonumber \\
\begin{Young} &\cr&\cr&\cr&\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1&1\cr&&2&2\cr&\cr&\cr \end{Young}\;,\; \begin{Young} &&1&1\cr&\cr&\cr&\cr2&2\cr \end{Young}\;,\;\begin{Young} &\cr&\cr&\cr&\cr 1&1\cr2 &2\cr\end{Young} \nonumber \\
\begin{Young} &\cr&\cr&\cr&\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0\\
\begin{Young}& &\cr&&\cr\cr \cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young}&&&1\cr&&&2\cr&1\cr&2 \cr \end{Young} \nonumber \\
\begin{Young}&& \cr&& \cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young}&&&1 \cr&&&2 \cr&3\cr&4\cr \end{Young} \;,\; \begin{Young}&& \cr&& \cr\cr\cr1\cr2\cr3\cr4 \cr\end{Young} \nonumber \\
\begin{Young}& \cr& \cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young}&&1 \cr&&2 \cr\cr\cr\cr\cr1\cr2\cr \end{Young}\nonumber
\end{eqnarray}
\\
\begin{eqnarray*}
\begin{Young}& \cr& \cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}&
\begin{Young}&&1 \cr&&2 \cr\cr\cr\cr\cr3\cr4\cr \end{Young}\;,\;\begin{Young}& \cr& \cr&1\cr&2\cr&3\cr&4\cr \end{Young} \\
\begin{Young} \cr\cr\cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr\cr\cr\cr\cr \end{Young}\\
\begin{Young} \cr\cr\cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0
\end{eqnarray*}
\\
Adding up all the Young diagram labels, we indeed get 13 operators. The reader is invited to check that for $SO(4)$, there are 12 operators and for $SO(6)$, there are 9 operators.
\subsubsection{3 $Z$'s and 3 $Y$'s for $SO(8)$}
As a final example, we consider 3 $Z$'s and 3 $Y$'s for $SO(8)$. The partition function gives a total of 14 operators. For $r$, we have
\begin{equation}
\label{eq:47}
\begin{Young} &&\cr&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
and for $s$
\begin{equation}
\label{eq:48}
\begin{Young} &&\cr&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying, we find
\begin{eqnarray}
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young} 1&1&1\cr2&2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&&1&1&1\cr&&&2&2&2\cr \end{Young} \;,\; \begin{Young} &&&1&1&1\cr&&&2\cr2&2\cr \end{Young},\; \begin{Young} &&&1\cr&&&2\cr1&1\cr2&2\cr \end{Young} \nonumber \\
\begin{Young} &\cr&\cr\cr\cr \end{Young}\otimes \begin{Young} 1&1&1\cr2&2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1&1\cr&&2&2\cr&1\cr&2\cr \end{Young} \nonumber \\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young} 1&1&1\cr2&2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1 \cr&2&2\cr\cr\cr\cr\cr1\cr2\cr \end{Young} \nonumber \\
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young} 1&1\cr2&2\cr.\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&&1\cr&&&2\cr1&3\cr2&4\cr \end{Young}\\
\begin{Young} &\cr&\cr\cr\cr \end{Young}\otimes \begin{Young} 1&1\cr2&2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young}&&1&1\cr&&2&2\cr&3\cr&4\cr \end{Young}\;,\; \begin{Young} &&1&1\cr&\cr&2\cr&3\cr2&4\cr \end{Young}\;,\; \begin{Young} &&1\cr&&2\cr\cr\cr1\cr2\cr3\cr4\cr \end{Young} \;,\; \begin{Young} &\cr&\cr&1\cr&2\cr1&3\cr2&4\cr \end{Young} \nonumber
\end{eqnarray}
\\
\begin{eqnarray*}
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}\otimes \begin{Young} 1&1\cr2&2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr\cr\cr3\cr4\cr \end{Young}\\
\begin{Young} &&\cr&&\cr \end{Young}\otimes \begin{Young} 1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} &\mathrm{gives}& \begin{Young} &&\cr&&\cr1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} \\
\begin{Young} &\cr&\cr\cr\cr \end{Young}\otimes \begin{Young} 1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr\cr\cr3\cr4\cr5\cr6\cr \end{Young}\\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}\otimes \begin{Young} 1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} &\mathrm{gives}& \begin{Young} &1\cr&2\cr&3\cr&4\cr&5\cr&6\cr \end{Young}
\end{eqnarray*}
\\
Adding up all the Young diagram labels, we get our expected 14 restricted Schur polynomials. Again the reader is invited to check that for $SO(6)$ there are 10 restricted Schurs and for $SO(4)$, there are 12 restricted Schurs.
\section{Appendix 2. Two-point function examples}
In the following examples, we denote the eigenvector of $\hat{P}_{[S]}$ by $\hat{\ket{[S]}}$.
\subsection{$q=2$}
For $q=2$, we have
\begin{equation}
\label{eq:50}
R = \begin{Young}&\cr&\cr\end{Young}, \hspace{20pt} (r,s) = (\begin{Young}\cr\cr\end{Young},\begin{Young}\cr\cr\end{Young})
\end{equation}
\\
Firstly
\begin{equation}
\label{eq:51}
\ket{[\hat{S}]} = \frac{1}{2}\,\begin{Young}4&3\cr2&1\cr\end{Young} + \frac{\sqrt{3}}{2}\,\begin{Young}4&2\cr3&1\cr\end{Young}, \hspace{20pt} \ket{[A],[A]} = \begin{Young}4&2\cr3&1\cr\end{Young}
\end{equation}
\\
The restricted Schur is
\begin{equation}
\label{eq:52}
O_{R(r,s)} = 2\sqrt{3}\mathrm{Tr}(ZY)
\end{equation}
\\
Evaluating the Wick contractions, its two point function is
\begin{equation}
\label{eq:53}
\langle O_{R(r,s)}\overline{O}_{T(t,u)} \rangle = 12 \langle \mathrm{Tr}(ZY) \mathrm{Tr}(\overline{Z}\overline{Y}) \rangle = 24N(N-1)
\end{equation}
\\
Using (\ref{eq:TPFformula}), with $q=2, n=m=1$ and $d_{R}=2$, we get
\begin{equation}
\label{eq:54}
\langle O_{R(r,s)}\overline{O}_{T(t,u)} \rangle = 24N(N-1).
\end{equation}
\\
\subsection{$q=4$}
First consider
\begin{equation}
\label{eq:55}
R = \begin{Young} &&&\cr&&&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &\cr&\cr \end{Young},\begin{Young} &\cr&\cr \end{Young})
\end{equation}
\\
We caluculate
\begin{equation}
\label{eq:ShatforR44}
\ket{[\hat{S}]} = \frac{\sqrt{5}}{4} \begin{Young} 8&6&4&2\cr7&5&3&1\cr \end{Young} + \sqrt{\frac{5}{48}} \begin{Young} 8&7&4&2\cr6&5&3&1\cr \end{Young} + \frac{\sqrt{5}}{12} \begin{Young} 8&7&4&3\cr6&5&2&1\cr \end{Young} + \frac{2}{3} \begin{Young} 8&7&6&5\cr4&3&2&1\cr \end{Young} + \sqrt{\frac{5}{48}}\begin{Young} 8&6&4&3\cr7&5&2&1\cr \end{Young}
\end{equation}
\\
and
\begin{equation}
\label{eq:AAvecforR44}
\ket{[A],[A]} = \begin{Young} 8&6&4&2\cr7&5&3&1\cr \end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:56}
O_{R(r,s)} = \frac{2\sqrt{5}}{3}\Big( 2\mathrm{Tr}(ZY)^{2} + 2\mathrm{Tr}(ZYZY) + \mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2}) + 4\mathrm{Tr}(Z^{2}Y^{2}) \Big)
\end{equation}
\\
and, after evaluating all the Wick contractions, its two-point function is
\begin{equation}
\label{eq:57}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N+1)(N+2)
\end{equation}
\\
Equation (\ref{eq:TPFformula}), for $q=4, n=m=2, d_{R} = 14$ and taking only the weights of the odd columns, gives
\begin{equation}
\label{eq:58}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N+1)(N+2)
\end{equation}
\\
Next consider
\begin{equation}
\label{eq:59}
R = \begin{Young} &\cr&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &\cr&\cr \end{Young},\begin{Young} &\cr&\cr \end{Young})
\end{equation}
\\
The state $\ket{[\hat{S}]}$ was found to be
\begin{equation}
\label{eq:60}
\ket{[\hat{S}]} = \frac{3}{4} \begin{Young}8 &6\cr7&5\cr4&2\cr3&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young}8 &7\cr6&5\cr4&2\cr3&1\cr \end{Young} + \frac{1}{4} \begin{Young}8 &7\cr6&5\cr4&3\cr2&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young} 8&6\cr7&5\cr4&3\cr2&1\cr \end{Young}
\end{equation}
\\
and
\begin{equation}
\label{eq:61}
\ket{[A],[A]} = \begin{Young}8 &6\cr7&5\cr4&2\cr3&1\cr \end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:62}
O_{R(r,s)} = \frac{2}{3}\Big( 2\mathrm{Tr}(ZY)^{2} + 2\mathrm{Tr}(ZYZY) + 3\mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2}) - 12\mathrm{Tr}(Z^{2}Y^{2}) \Big)
\end{equation}
\\
and its two-point function is
\begin{equation}
\label{eq:63}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
Equation (\ref{eq:TPFformula}), for $q=4, n=m=2, d_{R} = 14$ and taking the weights from the odd columns, gives
\begin{equation}
\label{eq:64}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
Next consider
\begin{equation}
\label{eq:65}
R = \begin{Young} &\cr&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} \cr\cr\cr\cr \end{Young},\begin{Young} \cr\cr\cr\cr \end{Young})
\end{equation}
\\
We find
\begin{equation}
\label{eq:ShatforR2222}
\ket{[\hat{S}]} = \frac{3}{4} \begin{Young}8 &6\cr7&5\cr4&2\cr3&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young}8 &7\cr6&5\cr4&2\cr3&1\cr \end{Young} + \frac{1}{4} \begin{Young}8 &7\cr6&5\cr4&3\cr2&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young} 8&6\cr7&5\cr4&3\cr2&1\cr \end{Young}
\end{equation}
\\
and
\begin{equation}
\label{eq:66}
\ket{[A],[A]} = \begin{Young}8 &4\cr7&3\cr6&2\cr5&1\cr \end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:67}
O_{R(r,s)} = \frac{4\sqrt{5}}{3}\Big( \mathrm{Tr}(ZY)^{2} - 2\mathrm{Tr}(ZYZY) \Big)
\end{equation}
\\
and its two-point function is
\begin{equation}
\label{eq:68}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
Equation (\ref{eq:TPFformula}) gives
\begin{equation}
\label{eq:69}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
For the final example for which $n=m=2$, consider
\begin{equation}
\label{eq:6911}
R = \begin{Young} &&&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &\cr&\cr \end{Young},\begin{Young} &\cr&\cr \end{Young})
\end{equation}
\\
We find the following for $\ket{[\hat{S}]}$ and $\ket{[A],[A]}$
\begin{eqnarray}
\ket{[\hat{S}]} &=& \frac{1}{4\sqrt{6}}\begin{Young} 8&7&2&1\cr6&5\cr4&3\cr \end{Young} + \frac{1}{4\sqrt{2}}\begin{Young} 8&6&2&1\cr7&5\cr4&3\cr \end{Young} + \frac{1}{8\sqrt{3}}\begin{Young} 8&7&3&1\cr6&5\cr4&2\cr \end{Young} + \frac{1}{8}\begin{Young} 8&6&3&1\cr7&5\cr4&2\cr \end{Young} + \frac{\sqrt{5}}{8}\begin{Young} 8&7&4&1\cr6&5\cr3&2\cr \end{Young} \nonumber \\
&&+\; \frac{\sqrt{15}}{8}\begin{Young} 8&6&4&1\cr7&5\cr3&2\cr \end{Young} - \frac{1}{12\sqrt{2}}\begin{Young} 8&7&4&3\cr6&5\cr2&1\cr \end{Young} - \frac{1}{4\sqrt{6}}\begin{Young} 8&6&4&3\cr7&5\cr2&1\cr \end{Young} - \frac{1}{8\sqrt{3}}\begin{Young} 8&7&4&2\cr6&5\cr3&1\cr \end{Young} - \frac{1}{8}\begin{Young} 8&6&4&2\cr7&5\cr3&1\cr \end{Young} \nonumber \\
&&+\;\frac{\sqrt{5}}{8}\begin{Young} 8&7&3&2\cr6&5\cr4&1\cr \end{Young} + \frac{\sqrt{15}}{8}\begin{Young} 8&6&3&2\cr7&5\cr4&1\cr \end{Young} + \frac{1}{6}\sqrt{\frac{5}{2}}\begin{Young} 8&7&6&5\cr4&3\cr2&1\cr \end{Young} + \frac{1}{2}\sqrt{\frac{5}{6}}\begin{Young} 8&7&6&5\cr4&2\cr3&1\cr \end{Young}
\end{eqnarray}
\\
and
\begin{equation}
\label{eq:70}
\ket{[A],[A]} = -\frac{3}{8}\begin{Young} 8&6&3&1\cr7&5\cr4&2\cr \end{Young} + \frac{\sqrt{15}}{8}\begin{Young} 8&6&4&1\cr7&5\cr3&2\cr \end{Young} - \frac{5}{8}\begin{Young} 8&6&4&2\cr7&5\cr3&1\cr \end{Young} + \frac{\sqrt{15}}{8}\begin{Young} 8&6&3&2\cr7&5\cr4&1\cr\end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:71}
O_{R(r,s)} = \frac{4}{3}\Big( -\mathrm{Tr}(ZY)^{2} - \mathrm{Tr}(ZYZY) + \mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2}) + \mathrm{Tr}(Z^{2}Y^{2}) \Big)
\end{equation}
\\
and its two-point function, after evaluating all the Wick contractions, is
\begin{equation}
\label{eq:72}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{160}{3}N(N-1)(N-2)(N+2)
\end{equation}
\\
Equation (\ref{eq:TPFformula}), for $q=4, n=m=2, d_{R} = 56$ and odd columns, gives
\begin{equation}
\label{eq:73}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{160}{3}N(N-1)(N-2)(N+2)
\end{equation}
\\
Now let's try two examples where $n=3, m=1$, i.e., the number of $Z$'s is not the equal to the number of $Y$'s. Consider
\begin{equation}
\label{eq:74}
R = \begin{Young} &&&\cr&&&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &&\cr&&\cr \end{Young},\begin{Young} \cr\cr \end{Young})
\end{equation}
\\
States $\hat{\ket{[S]}}$ and $\ket{[A],[A]}$ are still the same as in (\ref{eq:ShatforR44}) and (\ref{eq:AAvecforR44}). The operator is
\begin{equation}
\label{eq:75}
O_{R(r,s)} = \frac{4}{\sqrt{5}}\Big( \mathrm{Tr}(ZY)\mathrm{Tr}(Z^{2}) + 2 \mathrm{Tr}(Z^{3}Y) \Big)
\end{equation}
\\
and its two-point function is
\begin{equation}
\label{eq:76}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{256}{5}N(N-1)(N+1)(N+2).
\end{equation}
\\
Formula (\ref{eq:TPFformula}) reproduces exactly this result. As the final example, we consider
\begin{equation}
\label{eq:77}
R = \begin{Young} &\cr&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = ( \begin{Young} &\cr&\cr\cr\cr \end{Young}, \begin{Young} \cr\cr \end{Young})
\end{equation}
\\
The $\hat{\ket{[S]}}$ is the same as in (\ref{eq:ShatforR2222}). However
\begin{equation}
\label{eq:78}
\ket{[A],[A]} = \frac{\sqrt{5}}{3}\begin{Young} 8&4\cr7&3\cr6&2\cr5&1\cr\end{Young} + \frac{2}{3}\begin{Young} 8&6\cr7&5\cr4&2\cr3&1\cr\end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:79}
O_{R(r,s)} = \frac{4}{\sqrt{5}}\Big( \mathrm{Tr}(ZY)\mathrm{Tr}(Z^{2}) - 2 \mathrm{Tr}(Z^{3}Y) \Big)
\end{equation}
\\
The two-point function is
\begin{equation}
\label{eq:80}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{256}{5}N(N-1)(N-2)(N-3).
\end{equation}
\\
which, again, is exactly reproduced by (\ref{eq:TPFformula}).
\bibliographystyle{unsrt}
\section{}
\tableofcontents
\section{Introduction}
The most studied example of AdS/CFT is $\mathcal{N}=4$ super Yang-Mills theory with $U(N)$ gauge group, and its dual, type IIB string theory on AdS$_{5}\times S^{5}$. Testing the AdS/CFT correspondence in non-planar limits of the gauge theory and its dual string theory is a very interesting problem. In this regard, integrability would be a powerful tool. Therefore, it is interesting to investigate if the integrability, that was found in the planar limit \cite{Beisert}, is also present in non-planar limits. Furthermore, testing AdS/CFT in less supersymmetric settings is also an interesting problem. In particular we have the $1/4$-BPS sector of the theory in mind. The study of restricted Schur polynomials has yielded progress in both these directions. A restricted Schur is a local operator in the gauge theory which can be built from a variety of fields. For instance, we can use the complex scalar Higgs fields, the fermion fields as well as the gauge fields to build these operators. We consider restricted Schurs built using two types of complex scalar fields, $Z$ and $Y$ say. In this case, the definition is
\begin{equation}
\label{eq:RSUNDef}
\chi_{R(r,s)\alpha\beta}(Z,Y) = \frac{1}{n!m!}\sum\limits_{\sigma \in S_{n+m}} \mathrm{Tr}\big( P_{R\rightarrow (r,s)\alpha\beta} \Gamma_{R}(\sigma) \big) \mathrm{Tr}\big( \sigma Z^{\otimes n}\otimes Y^{\otimes m} \big)
\end{equation}
\\
In (\ref{eq:RSUNDef}), $R$ is a Young diagram with $n+m$ boxes, corresponding to an irreducible representation (irrep) of $S_{n+m}$ and $(r,s)$ are a pair of Young diagrams with $n$ and $m$ boxes respectively corresponding to an irrep of $S_{n}\times S_{m}$. $P_{R\rightarrow (r,s)\alpha\beta}$ is a projector in the labels $R$ and $(r,s)$ and an intertwiner in the labels $\alpha$ and $\beta$. These labels are the multiplicity labels with which irrep $(r,s)$ is subduced from $R$ when restricting from $S_{n+m}$ to $S_{n}\times S_{m}$. Further details of (\ref{eq:RSUNDef}) can be found in \cite{GGWSA1}, \cite{GGO} for example. There are many good reasons to study the operators defined in (\ref{eq:RSUNDef}) as we now explain.
The counting of states given by the partition function of the $1/4$-BPS sector of the free $U(N)$ theory was shown to match the counting of restricted Schurs that can be defined \cite{Collins}. Essentially, the partition function was expressed in terms of Littlewood Richardson coefficients which count the number of restricted Schurs for a given set of Young diagram labels, $R,(r,s)$. This result was generalised to an arbitrary product of $U(N)$ gauge groups in \cite{Quiver}.
We are interested in operators whose bare dimension grows parametrically with $N$. For these operators, the large $N$ limit of correlation functions is not captured by summing only planar diagrams \cite{GGO}. The usual $1/N^{2}$ factor suppressing non-planar diagrams is over-powered by combinatorial factors resulting from evaluating all possible Wick contractions. Indeed, when the number of matrix fields in the operator scales with $N$, we are forced to sum an infinite number of non-planar diagrams in the large $N$ limit. A solution to this problem was given in \cite{Jevicki} for the $1/2$-BPS case. In this important and influential work, it was observed that there is a one-to-one correspondence between the space of $1/2$-BPS representations and Young diagrams, i.e., Schur polynomials. Using representation theory of the symmetric and unitary group, the two-point function of Schur polynomials was computed \emph{exactly} in the free theory limit. \cite{EMC} then achieved the remarkable result of computing the two-point function of restricted Schurs (\ref{eq:RSUNDef}) exactly in the free theory. The operators diagonalised the two-point function and a very simple formula for the final result was given. Thus, the restricted Schurs provide an exactly orthogonal basis for the $1/4$-BPS sector of the free $U(N)$ gauge theory.
Apart from the interesting properties of (\ref{eq:RSUNDef}) as gauge theory objects, restricted Schurs also have an AdS/CFT dual. When the scaling dimension of operators grows with $N$, $\chi_{R(r,s)\alpha\beta}(Z,Y)$, with $R$ having long columns or rows ($O(N)$ boxes), has been argued to be dual to a system of excited giant gravitons \cite{Invasion}, \cite{GGO}. Concretely, such a system is realised as a system of giant gravitons with strings attached \cite{GGWSA1}. Thus, restricted Schurs are a useful tool for studying non-perturbative physics of both the gauge theory and its dual string theory.
In a similar approach, \cite{Kimura:2007wy} wrote down an orthogonal basis of gauge invariant operators using $Z$ and $Z^{\dagger}$ using Brauer algebras and their permutation sub-algebras in the free theory limit. \cite{Ram} then also studied $1/4$ and $1/8$-BPS operators at zero coupling, $g^{2}_{YM} = 0$. Also by making heavy use of representation theory, the multi-matrix multi-trace operators constructed here were argued to form a basis and to diagonalise their two-point function. When the operators in \cite{Ram} consist of two matrix fields $Z$ and $Y$ say, they were shown to be identical to the basis of restricted Schur polynomials \cite{Collins}. Progress has also been made in the counting and constructing of $1/4$ and $1/8$-BPS operators at weak coupling \cite{Weakcoup}.
Tremendous progress has already been achieved in the study of the anomalous dimension spectrum of restricted Schur polynomials; see \cite{GGO}, \cite{NPI}, \cite{Spring}, \cite{twoloop}, \cite{HigherLoopAnDim} for examples. Anomalous dimensions of the restricted Schurs are dual to excitation energies of the excited giant graviton system. One of the main results of these works is that the spectrum of anomalous dimensions reduces to the set of normal mode frequencies of a system of decoupled harmonic oscillators. The fact that we get decoupled oscillators is evidence of integrability in the non-planar sectors studied in these works.
Is there a similar story for the theory with an $SO(N)$ gauge group? In this work, we take the first steps toward answering this question. $\mathcal{N} = 4$ super Yang-Mills theory with an $SO(N)$ gauge group is dual to type IIB string theory with AdS$_{5}\times \mathcal{R}$P$^{5}$ geometry. At the non-planar level, the spectral problems of $U(N)$ and $SO(N)$ gauge theories are distinct \cite{SpecProbSON}. Thus, there is a good chance that a study of anomalous dimensions of large dimension operators will reveal new aspects of D-brane physics. The $1/2$-BPS case for free $SO(N)$ gauge theory was studied in \cite{SON1}, \cite{SON2} and \cite{Diaz} . The results of \cite{SON1} and \cite{SON2} include defining Schur polynomials in the square of the eigenvalues and computing their two-point functions exactly in the free theory. Here too, the Schur polynomial was shown to diagonalise the two-point function, a property that was argued to be independent of the gauge group \cite{Diaz}.
We show that restricted Schur polynomials provide an exactly orthogonal basis for the $1/4$-BPS sector of free super Yang-Mills theory with $SO(N)$ gauge group. In this work, we take $N$ to be even. First we show that the counting of states obtained from the partition function in this sector matches the number of restricted Schurs that can be defined. This is achieved by expressing the partition function in terms of the Littlewood Richardson coefficients. The counting comes out differently from the $1/4$-BPS $U(N)$ case as we will see. $SO(N)$ has two invariant tensors. They are $\delta_{ij}$ and $\epsilon_{i_{1}i_{2}\cdots i_{N}}$. We focus on operators constructed using $\delta_{ij}$. We construct restricted Schur polynomials and manage to compute the two-point function exactly in the free theory. The result is expressed in a relatively simple formula. In the next section, we present some background theory needed for our computations.
\section{Projectors}
In this section, we briefly discuss some of the projectors and states that appear frequently in our calculations. All projectors we discuss here project from the carrier space of irrep $R$ of $S_{2q}$, $q=n+m$, onto the carrier space of an irrep of some subgroup.
\subsection{From $R$ onto $(r,s)$}
Firstly we discuss the projectors $P_{R\rightarrow (r,s)\alpha\beta}$ constructed in \cite{GGO}. They project from $R$ onto irrep $(r,s)$ of the subgroup $S_{2n}\times S_{2m}$. In the multiplicity labels, however, $P_{R\rightarrow (r,s)\alpha\beta}$ is an intertwiner \cite{RepTh}. We write these operators as
\begin{equation}
\label{eq:1}
P_{R\rightarrow (r,s)\alpha\beta} = \sum\limits^{d_{r}\times d_{s}}_{l=1} \ket{R(r,s)\alpha;l}\!\!\bra{R(r,s)\beta;l}
\end{equation}
\\
where $\ket{R(r,s)\alpha;l}$ is a state in the carrier space of the $\alpha$-th copy of irrep $(r,s)$ and $d_{r}\times d_{s}$ is the dimension of $(r,s)$. In this basis, the matrix representation of $\gamma \in S_{2n}\times S_{2m}$ in irrep $R$ is block diagonal with the $S_{2n}\times S_{2m}$ irreps as the diagonal blocks. For example
\begin{equation}
\label{eq:2}
\Gamma_{R}(\gamma) = \left(\begin{array}{cccc}\Gamma_{(r',s')}(\gamma) & 0 & 0 & 0 \\0 & \Gamma_{(r'',s'')}(\gamma) & 0 & 0 \\0 & 0 & \Gamma_{(r,s)1}(\gamma) & 0 \\0 & 0 & 0 & \Gamma_{(r,s)2}(\gamma)\end{array}\right)
\end{equation}
\\
where we have shown two copies of irrep $(r,s)$. There are four projectors that can be defined for $(r,s)$. Acting on $\Gamma_{R}(\gamma)$ each give the following
\begin{eqnarray}
P_{R\rightarrow(r,s)11}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & \Gamma_{(r,s)1}(\gamma) & 0 \\0 & 0 & 0 & 0\end{array}\right)\\
P_{R\rightarrow(r,s)22}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & \Gamma_{(r,s)2}(\gamma)\end{array}\right)\\
P_{R\rightarrow(r,s)12}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & \Gamma_{(r,s)2}(\gamma) \\0 & 0 & 0 & 0\end{array}\right)\\
P_{R\rightarrow(r,s)21}\Gamma_{R}(\gamma) &=& \left(\begin{array}{cccc}0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 \\0 & 0 & \Gamma_{(r,s)1}(\gamma) & 0\end{array}\right)
\end{eqnarray}
\\
\subsection{From $R$ onto $[S]$}
Let $R$ have an even number of boxes in each row. Restricting $S_{2q}$ to the subgroup $S_{q}[S_{2}]$, one can find a basis in which $\Gamma_{R}(\xi)$, $\xi\in S_{q}[S_{2}]$, is block diagonal with the $S_{q}[S_{2}]$ irreps appearing on the diagonal. For $R$ having even rows, there exists a 1-dimensional irrep of $S_{q}[S_{2}]$ which we denote $[S]$. In this irrep, all matrices are represented by 1. Let $\ket{[S]}$ be the state spanning the carrier space of $[S]$. Then we have \cite{Ivanov}
\begin{equation}
\label{eq:3}
\Gamma_{R}(\xi)\ket{[S]} = \ket{[S]}, \hspace{20pt} \xi \in S_{q}[S_{2}].
\end{equation}
\\
We can also define a projector to go from $R$ to $[S]$.
\begin{equation}
\label{eq:4}
P_{[S]} = \frac{1}{q!2^{q}}\sum\limits_{\xi\in S_{q}[S_{2}]}\Gamma_{R}(\xi)
\end{equation}
\\
The state $\ket{[S]}$ may be calculated as the eigenvector of $P_{[S]}$ with eigenvalue 1. Lastly, $[S]$ is subduced from $R$ with multiplicity 1 \cite{SON1}.
\subsection{From $(r,s)\beta$ to $([A],[A])\beta$}
Recall that $(r,s)\beta$ is the $\beta$-th copy of the irrep of $S_{2n}\times S_{2m}$. If $r$ has an even number of boxes in each column, then, upon restricting to the $S_{n}[S_{2}]$ subgroup another type of 1 dimensional irrep may be subduced \cite{Ivanov}. Call this irrep $[A]$. In this irrep,
\begin{equation}
\label{eq:5}
\Gamma_{r}(\eta) = \mathrm{sgn}(\eta), \hspace{20pt} \eta \in S_{n}[S_{2}].
\end{equation}
\\
Denote by $\ket{[A]}$ the state spanning the 1-dimensional carrier space of $[A]$. Irrep $[A]$ is also subduced from $r$ with no multiplicity \cite{SON1}. If $r$ and $s$ both have even columns, then each may subduce the irrep $[A]$ of $S_{n}[S_{2}]$ and $S_{m}[S_{2}]$ respectively. Thus, $(r,s)\beta$ subduces the irrep $([A],[A])\beta$ of $S_{n}[S_{2}]\times S_{m}[S_{2}]$. Denote by $\ket{[A],[A]\beta}$ the state spanning this 1-dimensional irrep. Define
\begin{equation}
\label{eq:6}
P_{R\rightarrow ([A],[A])\beta} = P_{R\rightarrow(r,s)\beta\beta}P_{[A,A]}, \hspace{20pt} P_{[A,A]} = \frac{1}{n!m!2^{q}}\sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]}\mathrm{sgn}(\eta)\Gamma_{R}(\eta)
\end{equation}
\\
The state $\ket{[A],[A]\beta}$ may be defined to be the eigenvector of $P_{R\rightarrow ([A],[A])\beta}$ with eigenvalue 1. This state has the following properties
\begin{enumerate}
\item \begin{equation}
\label{eq:611}
P_{R\rightarrow (r,s)\alpha\beta} \ket{[A],[A]\beta} = \ket{[A],[A]\alpha}
\end{equation}
\item
\begin{equation}
\Gamma_{R}(\eta) \ket{[A],[A]\beta} = \mathrm{sgn}(\eta)\ket{[A],[A]\beta},\hspace{20pt} \eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]
\end{equation}
\item The quantity $\langle[A],[A]\beta | R(r,s)\beta;l\rangle $ does not depend on the multiplicity index $\beta$. This will be important when we construct the restricted Schurs.
\end{enumerate}
\section{Counting}
In this section we express the partition function for the $1/4$-BPS sector of free $SO(2n)$ gauge theory in terms of the Littlewood Richardson coefficients depending on three Young diagram labels. Two Young diagrams will have an even number of boxes in each \emph{column}, and the third one (induced from the previous two) will have an even number of boxes in each \emph{row}. This is first demonstrated for the $SO(4)$ case for simplicity. It is then easy to extend the argument to the general $SO(2n)$ case, which follows thereafter. The Littlewood Richardson number $g(\lambda,\mu,\xi)$ counts the number of times a Young diagram $\xi$ is induced from two smaller Young diagrams $\mu$ and $\lambda$, say. Thus, $g(\lambda,\mu,\xi)$ counts the number of restricted Schurs that may be defined for labels $\{\lambda,\mu,\xi\}$.
The partition function for the $1/4$-BPS sector of free $SO(2n)$ gauge theory is \cite{SON1}, \cite{Dolan1}
\begin{eqnarray}
\label{eq:SO2nPF}
G(t_{1},t_{2}) &=& \frac{1}{2^{n-1}n!}\int_{T_{n}} \prod\limits^{n}_{j=1} \frac{dx_{j}}{2\pi ix_{j} }\Delta(x+x^{-1})^{2}\times \\
&& \prod\limits^{2}_{k=1}\prod\limits_{1\leqslant i<j\leqslant n}\frac{1}{(1-t_{k})^{n}}\frac{1}{(1-t_{k}x_{i}x_{j})(1-t_{k}x_{i}x^{-1}_{j})(1-t_{k}x^{-1}_{i}x_{j})(1-t_{k}x^{-1}_{i}x^{-1}_{j})} \nonumber
\end{eqnarray}
\\
Let's study the simple $SO(4)$ case first and generalise to $SO(2n)$ thereafter. For $SO(4)$, this becomes
\begin{eqnarray}
\label{eq:1steqnforG}
G(t_{1},t_{2}) &=& \frac{1}{4}\int \frac{dx_{1}}{2\pi ix_{1}}\int \frac{dx_{2}}{2\pi ix_{2}}\Delta(x+x^{-1})^{2}\times \\
&& \prod\limits^{2}_{k=1}\frac{1}{(1-t_{k})^{2}}\frac{1}{(1-t_{k}x_{1}x_{2})(1-t_{k}x_{1}x^{-1}_{2})(1-t_{k}x^{-1}_{1}x_{2})(1-t_{k}x^{-1}_{1}x^{-1}_{2})}\nonumber
\end{eqnarray}
\\
The factor in the second line may be written as
\begin{eqnarray}
&&\frac{1}{(1-t_{k}x_{1}x^{-1}_{1})(1-t_{k}x_{2}x^{-1}_{2})(1-t_{k}x_{1}x_{2})(1-t_{k}x_{1}x^{-1}_{2})(1-t_{k}x^{-1}_{1}x_{2})(1-t_{k}x^{-1}_{1}x^{-1}_{2})} \nonumber \\
\label{eq:rewritingprod}
&&\hspace{200pt} =\; \prod\limits_{1\leqslant i< j \leqslant 4}\frac{1}{1-t_{k}y_{i}y_{j}}
\end{eqnarray}
\\
with $y_{1}=x_{1}, y_{2} = x^{-1}_{1}, y_{3} = x_{2}, y_{4} = x^{-1}_{2}$. Now we may expand this product in terms of Schur polynomials using the formula \cite{Ellitic}
\begin{equation}
\label{eq:modcaulittd}
\prod\limits_{1\leqslant i< j \leqslant L}\frac{1}{1-y_{i}y_{j}} = \sum\limits_{\mu}s_{\mu^{2}}(y_{1},..., y_{L})
\end{equation}
\\
where $\mu^{2}$ is defined as the partition with parts
\begin{equation}
\label{eq:7}
(\mu^{2})_{i} = \mu_{\lceil i/2 \rceil}.
\end{equation}
\\
This means the following. Take, for example, $i = \{1,2,3,4\}$. Then, $i/2 = \{ 1/2, 1, 3/2, 2 \}$. The ceiling symbol then tells us to take: $\lceil i/2 \rceil = \{ 1,1,2,2 \}$. So, if $\mu^{2}\,_{i}$ denotes the $i$th part of $\mu^{2}$, then
\begin{equation}
\label{eq:8}
\mu^{2} = (\mu^{2}\,_{1},\mu^{2}\,_{2},\mu^{2}\,_{3},\mu^{2}\,_{4}) = (\mu_{1},\mu_{1},\mu_{2},\mu_{2})
\end{equation}
\\
Thus, in general for a partition $\mu^{2}$ with $n$ parts,
\begin{equation}
\label{eq:9}
\mu^{2} = (\mu^{2}\,_{1},\mu^{2}\,_{2},\mu^{2}\,_{3},\mu^{2}\,_{4},\cdots, \mu^{2}\,_{n-1},\mu^{2}\,_{n}) = (\mu_{1},\mu_{1},\mu_{2},\mu_{2},\cdots,\mu_{n/2},\mu_{n/2})
\end{equation}
\\
Partitions $\mu^{2}$ then correspond to Young diagrams with an even number of boxes in each column. Including the prefactor $t_{k}$ in (\ref{eq:modcaulittd}), we get
\begin{equation}
\label{eq:911}
\prod\limits_{1\leqslant i< j \leqslant L}\frac{1}{1-t_{k}y_{i}y_{j}} = \sum\limits_{\mu}s_{\mu^{2}}(\sqrt{t_{k}}y_{1},..., \sqrt{t_{k}}y_{L}) = \sum\limits_{\mu} (t_{k})^{|\mu^{2}|/2} s_{\mu^{2}}(y_{1},...,y_{L})
\end{equation}
\\
where we used the fact that Schur polynomial $s_{\lambda}(y_{1},y_{2},\cdots, y_{L})$ may be written as
\begin{equation}
\label{eq:10}
s_{\lambda}(ty_{1},ty_{2},\cdots,ty_{L}) = t^{|\lambda|}s_{\lambda}(y_{1},y_{2},\cdots, y_{L}).
\end{equation}
\\
In the above formulas, $|\,|$ stands for the weight of the partition. The partition function (\ref{eq:1steqnforG}) becomes
\begin{eqnarray}
\label{eq:11}
G(t_{1},t_{2}) &=& \frac{1}{4}\int \frac{dx_{1}}{2\pi ix_{1}}\int \frac{dx_{2}}{2\pi ix_{2}}\Delta(x+x^{-1})^{2}\times \\
&&\sum\limits_{\lambda}\sum\limits_{\mu}(t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2}s_{\lambda^{2}}(x_{1},x^{-1}_{1},x_{2},x^{-1}_{2})s_{\mu^{2}}(x_{1},x^{-1}_{1},x_{2},x^{-1}_{2})\nonumber
\end{eqnarray}
\\
Now, we use the product rule involving Schur polynomials. This means that we can write the product of two Schur polynomials with partitions $\lambda^{2}$ and $\mu^{2}$ as the sum over a single Schur with partition $\xi$ times by the Littlewood Richardson coefficient $g(\lambda^{2},\mu^{2},\xi)$ \cite{Dolan2}. The Littlewood Richardson coefficient is how many times $\xi$ subduces $(\lambda^{2},\mu^{2})$. Thus,
\begin{eqnarray}
\label{eq:12}
G(t_{1},t_{2}) &=& \sum\limits_{\xi}\sum\limits_{\lambda}\sum\limits_{\mu} g(\lambda^{2},\mu^{2},\xi) (t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2} \times\\
&& \frac{1}{4}\int \frac{dx_{1}}{2\pi ix_{1}}\int \frac{dx_{2}}{2\pi ix_{2}}\Delta(x+x^{-1})^{2}s_{\xi}(x_{1},x^{-1}_{1},x_{2},x^{-1}_{2}) \nonumber
\end{eqnarray}
\\
Next, we generalise this to $SO(2n)$. In the same way as in equation (\ref{eq:rewritingprod})
\begin{equation}
\label{eq:13}
\frac{1}{(1-t_{k})^{n}}\prod\limits_{1\leqslant i <j \leqslant n}\frac{1}{(1-t_{k}x_{i}x_{j})(1-t_{k}x_{i}x^{-1}_{j})(1-t_{k}x^{-1}_{i}x_{j})(1-t_{k}x^{-1}_{i}x^{-1}_{j})} = \prod\limits_{1\leqslant i< j \leqslant N}\frac{1}{1-t_{k}y_{i}y_{j}}
\end{equation}
\\
with $y_{i} = x_{\lceil i/2 \rceil}$ and $y_{2i} = x^{-1}_{i}$ for $i = 1,2,\cdots, n$. Using equation (\ref{eq:modcaulittd}) for each factor,
\begin{equation}
\label{eq:14}
\prod\limits^{2}_{k=1} \prod\limits_{1\leqslant i< j \leqslant N}\frac{1}{1-t_{k}y_{i}y_{j}} = \sum\limits_{\lambda}\sum\limits_{\mu}(t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2}s_{\lambda^{2}}(y_{1},..,y_{N})s_{\mu^{2}}(y_{1},..,y_{N})
\end{equation}
\\
After using the product rule again, the $SO(2n)$ partition function in (\ref{eq:SO2nPF}) becomes
\begin{eqnarray}
\label{eq:InteformofPF}
G(t_{1},t_{2}) &=& \sum\limits_{\xi}\sum\limits_{\lambda}\sum\limits_{\mu} g(\lambda^{2},\mu^{2},\xi) (t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2} \times\\
&& \frac{1}{2^{n-1}n!}\int_{T_{n}} \prod\limits^{n}_{j=1} \frac{dx_{j}}{2\pi ix_{j}}\Delta(x+x^{-1})^{2}s_{\xi}(x_{1},x^{-1}_{1},...,x_{n},x^{-1}_{n})\nonumber
\end{eqnarray}
\\
Recognising the Haar (or $G$-invariant) measure for $SO(2n)$, the integral in (\ref{eq:InteformofPF}) may be written as
\begin{equation}
\label{eq:15}
I_{n} = \int_{SO(2n)}[dO] \,s_{\xi}(\mathrm{x})
\end{equation}
\\
This integral is equal to 1 for two cases. The first is if $\xi$ has an even number of boxes in each \emph{row}. The second is if $\xi$ has a Young diagram with an even number of boxes in each row stuck onto a single column of $2n$ boxes\footnote{Recall that there are only two invariant tensors for $SO(2n)$. They are $\delta_{ij}$ and $\epsilon_{i_{1}i_{2}\cdots i_{2n}}$. This case is relevent for operators that are built using $\epsilon_{i_{1}i_{2}\cdots i_{2n}}$.}. Here are examples of each for $SO(4)$
\begin{equation}
\label{eq:16}
\xi = \begin{Young} &&&\cr&\cr&\cr \end{Young}, \hspace{20pt} \mathrm{or} \hspace{20pt} \xi = \begin{Young} &&\cr&&\cr\cr\cr \end{Young}
\end{equation}
\\
The integral vanishes in all other cases. See \cite{Ellitic} \cite{Vanishing} and \cite{SymmFunc}. We have checked these results for many examples for $SO(4)$ and $SO(6)$. With this result, the $SO(2n)$ partition function becomes
\begin{equation}
\label{eq:FinalPF}
G(t_{1},t_{2}) = \sum\limits_{\xi}\sum\limits_{\lambda}\sum\limits_{\mu} g(\lambda^{2},\mu^{2},\xi) (t_{1})^{|\lambda^{2}|/2}(t_{2})^{|\mu^{2}|/2}
\end{equation}
\\
where $\xi$ is a Young diagram with even rows and $\lambda^{2}$ and $\mu^{2}$ are Young diagrams with even columns. This result is truly different from the $U(N)$ $1/4$-BPS case, studied in \cite{Collins}. For $U(N)$ the counting went according to the square of the Littlewood-Richardson number. Here, the counting goes according to the Littlewood Richardson number itself. In other words, for a given $S_{2n}\times S_{2m}$ irrep, $(r,s)$, each with even columns, and given $S_{2q}$ irrep $R$ with even rows, the number of restricted Schurs one can define is exactly $g(r,s,R)$. We present some counting examples in appendix 1.
\section{Defining the restricted Schurs}
We now discuss the problem of defining the restricted Schurs for the $1/4$-BPS free $SO(N)$ gauge theory. We focus on operators that can be constructed using $\delta$ rather than $\epsilon_{i_{1}i_{2}\cdots i_{2n}}$. We construct these operators out of two complex matrix fields $Z$ and $Y$. The two-point function of these fields is given by \cite{SON1}
\begin{equation}
\label{eq:17}
\langle Z^{ij}\overline{Z}^{kl} \rangle = \langle Z^{ij}Z_{kl} \rangle = \delta^{i}_{k}\delta^{j}_{l} - \delta^{i}_{l}\delta^{j}_{k}
\end{equation}
\\
and similarly for $Y$. We define our $SO(N)$ restricted Schurs to be
\begin{equation}
\label{eq:RSDef}
O_{R(r,s)\alpha}(Z,Y) = \frac{1}{(2n)!(2m)!}\sum\limits_{\sigma \in S_{2q}} \mathrm{Tr}\big( \mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)C^{4\nu}_{I}\sigma^{I}_{J} (Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}
\end{equation}
\\
where
\begin{equation}
\label{eq:ODef}
\mathcal{O}_{R(r,s)\alpha} = \ket{[S]}\!\!\bra{[A],[A]\beta}P_{R\rightarrow (r,s)\beta\alpha}
\end{equation}
\\
In (\ref{eq:RSDef}), the tensor $(Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}$ is defined as
\begin{equation}
\label{eq:18}
(Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J} = Z^{j_{1}j_{2}}\cdots Z^{j_{2n-1}j_{2n}}Y^{j_{2n+1}j_{2n+2}}\cdots Y^{j_{2q-1}j_{2q}}
\end{equation}
\\
$C^{4\nu}_{I}$ is the contractor defined in \cite{SON2}
\begin{equation}
\label{eq:19}
C^{4\nu}_{I} = \delta_{k_{1}k_{2}}\cdots \delta_{k_{2q-1}k_{2q}}(\sigma_{4\nu})^{K}_{I}
\end{equation}
\\
$C^{4\nu}_{I}$ is responsible for contracting the free indices in such a way to make the operator gauge invariant. For $\sigma_{4\nu}$, we take
\begin{equation}
\label{eq:20}
\sigma_{4\nu} = (1,2,3,4)(5,6,7,8)\cdots (2q-3,2q-2,2q-1,2q)
\end{equation}
\\
This permutation contracts indices $1\&4$, $2\&3$ $\cdots$, $2q-3\&2q$ and $2q-2\&2q-1$, i.e.,
\begin{equation}
\label{eq:21}
C^{4\nu}_{I}(\sigma)^{I}_{J} = \sigma^{i_{1}i_{2}i_{2}i_{1}\cdots i_{q-1}i_{q}i_{q}i_{q-1}}_{j_{1}j_{2}j_{3}j_{4}\cdots j_{2q}}.
\end{equation}
\\
$R$ is an irrep of $S_{2q}$ and $(r,s)$ is an irrep of the subgroup $S_{2n}\times S_{2m}$. Indices $\alpha$ and $\beta$ are multiplicity labels for the irrep $(r,s)$. State $\ket{[A],[A]\beta}$ is the state spanning the one-dimensional carrier space $([A],[A])\beta$ of $S_{n}[S_{2}]\times S_{m}[S_{2}]$ as defined in section 2. Concretely, it is the eigenvector of the operator $P_{R\rightarrow ([A],[A])\beta}$ with eigenvalue 1. State $\ket{[S]}$ is the state spanning the 1-dimensional carrier space $[S]$ of $S_{q}[S_{2}]$ as defined in section 2. The $S_{q}[S_{2}]$ here is chosen to stabilise the set $(1,4),(2,3),\cdots, (2q-3,2q),(2q-2,2q-1)$. Concretely, $\ket{[S]}$ is the eigenvector of $P_{[S]}$.
The coefficients in (\ref{eq:RSDef}) satisfy the following property
\begin{equation}
\label{eq:22}
\mathrm{Tr}\big( \mathcal{O}_{R(r,s)\alpha} \Gamma_{R}(\sigma) \big)\mathrm{sgn}(\eta) = \mathrm{Tr}\big( \mathcal{O}_{R(r,s)\alpha} \Gamma_{R}(\eta\sigma\xi) \big), \hspace{20pt} \xi \in S_{q}[S_{2}], \eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]
\end{equation}
\\
Lastly, the operators in (\ref{eq:RSDef}) depend only on a single multiplicity label and thus, match the counting found in section 3.
\section{Two-point function}
In this section, we evaluate the two-point function of the operators defined in section 4. Recall the definition for our restricted Schurs,
\begin{eqnarray}
O_{R(r,s)\alpha} &=& \frac{1}{(2n)!(2m)!}\sum\limits_{\sigma\in S_{2q}}\mathrm{Tr}(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R})C^{4\nu}_{I}\sigma^{I}_{J}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}\\
\overline{O}_{T(t,u)\beta} &=& \frac{1}{(2n)!(2m)!}\sum\limits_{\sigma\in S_{2q}}\mathrm{Tr}(\mathcal{O}_{T(t,u)\beta}\Gamma_{R})C^{K}_{4\nu}(\overline{\sigma})^{L}_{K}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})_{L}
\end{eqnarray}
\\
where
\begin{equation}
\label{eq:23}
(\overline{\sigma})^{J}_{I} = \delta^{j_{\sigma(1)}}_{i_{1}}\delta^{j_{\sigma(2)}}_{i_{2}}\cdots \delta^{j_{\sigma(2q)}}_{i_{2q}} = \delta^{j_{1}}_{i_{\sigma^{-1}(1)}}\delta^{j_{2}}_{i_{\sigma^{-1}(2)}}\cdots \delta^{j_{2q}}_{i_{\sigma^{-1}(2q)}} = (\sigma^{-1})^{J}_{I}.
\end{equation}
\\
The first step in the calculation is evaluating the correlator
\begin{equation}
\label{eq:CorrZY}
\langle (Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})_{L} \rangle
\end{equation}
\\
In \cite{SON1} the correlator $(\langle Z^{\otimes 2n})^{J}(Z^{\otimes 2n})_{L} \rangle$ was found to be a sum over permutations belonging to the wreath product $S_{n}[S_{2}]$, where each term in the sum was weighted by the $\mathrm{sgn}$ of the permutation. The correlator in (\ref{eq:CorrZY}) generalises to a sum over the subgroup $S_{n}[S_{2}]\times S_{m}[S_{2}]$
\begin{equation}
\label{CorrZYanswer}
\langle (Z^{\otimes 2n} \otimes Y^{\otimes 2m})^{J}(Z^{\otimes 2n} \otimes Y^{\otimes 2m})_{L} = \sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]} \mathrm{sgn}(\eta) \eta^{J}_{L}
\end{equation}
\\
This formula may be easily checked for the example of $n=m=2$.
\begin{eqnarray}
&&\langle Z^{i_{1}i_{2}}Z^{i_{3}i_{4}}Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Z}^{j_{1}j_{2}}\bar{Z}^{j_{3}j_{4}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle = \nonumber \\
&&\langle Z^{i_{1}i_{2}}\bar{Z}^{j_{1}j_{2}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{3}j_{3}} \rangle \Big[ \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{5}j_{6}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{7}j_{8}} \rangle + \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{7}j_{8}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}} \rangle \Big] +\nonumber\\
&& \langle Z^{i_{1}i_{2}}\bar{Z}^{j_{3}j_{4}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{1}j_{2}} \rangle \Big[ \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{5}j_{6}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{7}j_{8}} \rangle + \langle Y^{i_{5}i_{6}}\bar{Y}^{j_{7}j_{8}} \rangle\langle Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}} \rangle \Big]\nonumber\\
&=& \langle Z^{i_{1}i_{2}}\bar{Z}^{j_{1}j_{2}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{3}j_{3}} \rangle \langle Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle + \langle Z^{i_{1}i_{2}}\bar{Z}^{j_{3}j_{4}} \rangle\langle Z^{i_{3}i_{4}}\bar{Z}^{j_{1}j_{2}} \rangle \langle Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle \nonumber\\
&=& \langle Z^{i_{1}i_{2}}Z^{i_{3}i_{4}}\bar{Z}^{j_{1}j_{2}}\bar{Z}^{j_{3}j_{4}} \rangle \langle Y^{i_{5}i_{6}}Y^{i_{7}i_{8}}\bar{Y}^{j_{5}j_{6}}\bar{Y}^{j_{7}j_{8}} \rangle \nonumber\\
&=& \bigg[\sum\limits_{\rho\in S_{2}[S_{2}]}\mathrm{Sgn}(\rho)\rho\bigg] \bigg[\sum\limits_{\sigma\in S_{2}[S_{2}]}\mathrm{Sgn}(\sigma)\sigma\bigg]\nonumber\\
&=& \sum\limits_{\psi \in S_{2}[S_{2}]\times S_{2}[S_{2}]}\mathrm{Sgn}(\psi)\psi^{I}_{J}
\end{eqnarray}
\\
where $\psi^{I}_{J} = \rho \sigma$ and $\mathrm{Sgn}(\psi)=\mathrm{Sgn}(\rho)\mathrm{Sgn}(\sigma)$. Now compute the two-point function
\begin{eqnarray}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle &=& \frac{1}{\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\sigma,\rho\in S_{2q}}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)\mathrm{Tr}\big(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho)\big)C^{4\nu}_{I}C^{L}_{4\nu}\sigma^{I}_{J}(\rho^{-1})^{K}_{L} \nonumber \\
&& \times \langle (ZY)^{J}(ZY)_{K} \rangle \nonumber \\
&=& \frac{1}{\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\sigma,\rho\in S_{2q}}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)\mathrm{Tr}\big(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho)\big)C^{4\nu}_{I}C^{L}_{4\nu}\sigma^{I}_{J}(\rho^{-1})^{K}_{L} \nonumber \\
&& \sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]}\mathrm{sgn}(\eta)\eta^{J}_{K} \nonumber \\
&=& \frac{1}{\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\sigma,\rho\in S_{2q}}\sum\limits_{\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\sigma)\big)\mathrm{Tr}\big(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho)\big)\mathrm{sgn}(\eta) \nonumber \\
\label{eq:TPF1}
&& \hspace{50pt} \times \;C^{4\nu}_{I}C^{L}_{4\nu}(\rho^{-1}\eta\sigma)^{I}_{L}
\end{eqnarray}
\\
Relabelling the sum over $\sigma = \eta^{-1}\psi$ where $\psi \in S_{2q}$, we obtain
\\
\begin{equation}
\label{eq:TPF2}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \frac{n!m!2^{q}}{\big( (2n)!(2m)! \big)^{2}} \sum\limits_{\psi,\rho\in S_{2q}}\mathrm{Tr}\big(\mathcal{O}_{R(r,s)\alpha}\Gamma_{R}(\psi)\big)\mathrm{Tr}(\mathcal{O}_{T(t,u)\beta}\Gamma_{T}(\rho))C^{4\nu}_{I}C^{L}_{4\nu}(\rho^{-1}\psi)^{I}_{L}
\end{equation}
\\
We also used the fact that $\eta \in S_{n}[S_{2}]\times S_{m}[S_{2}]$, which is a subgroup of $S_{2n}\times S_{2m}$ and that
\begin{equation}
\label{eq:24}
\bra{[A],[A]\gamma}\Gamma_{R}(\eta^{-1}) = \bra{[A],[A]\gamma}\mathrm{sgn}(\eta^{-1})
\end{equation}
\\
which then canceled with $\mathrm{sgn}(\eta)$ in (\ref{eq:TPF1}). Next, relabel the sum over $\psi$ by letting $ \psi = \rho\tau$, with $\tau \in S_{2q}$. After using the orthogonality relation
\begin{equation}
\label{eq:25}
\sum\limits_{\rho \in S_{2q}} \Gamma_{R}(\rho)_{ij}\Gamma_{T}(\rho)_{kl} = \frac{(2q)!}{d_{R}}\delta_{RT}\delta_{ik}\delta_{jl}
\end{equation}
\\
and using
\begin{equation}
\label{eq:26}
\mathcal{O}_{R(r,s)\alpha}\mathcal{O}^{T}_{T(t,u)\beta} = \ket{[S]}\!\!\bra{[A],[A]\gamma}P_{R\rightarrow(r,s)\gamma\zeta} \ket{[A],[A]\zeta}\!\!\bra{[S]}\delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta},
\end{equation}
\\
where $T$ in the superscript is the transpose of $\mathcal{O}$, equation (\ref{eq:TPF2}) becomes
\begin{equation}
\label{eq:TPF3}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\tau\in S_{2q}}\mathrm{Tr}\big(\hat{P}_{[S]}\Gamma_{R}(\tau)\big)C^{4\nu}_{I}C^{L}_{4\nu}(\tau)^{I}_{L}
\end{equation}
\\
where we also used the fact that
\begin{equation}
\label{eq:27}
\bra{[A],[A]\gamma}P_{R\rightarrow(r,s)\gamma\zeta}\ket{[A],[A],\zeta} = 1
\end{equation}
\\
to simplify the two-point function. In (\ref{eq:TPF3}), the $S_{q}[S_{2}]$ in $\hat{P}_{[S]}$ is the stabiliser for $(1,4)(2,3)\cdots$. For reasons that will become clear shortly, we want to change the embedding of the $S_{q}[S_{2}]$. Let $\rho$ be a permutation such that
\begin{equation}
\label{eq:28}
C^{4\nu}_{I}(\rho)^{I}_{K} = C^{4\mu}_{K}.
\end{equation}
\\
where $\sigma_{4\mu} = (1,2)(3,4) \cdots (2q-1,2q)$. Defining
\begin{equation}
\label{eq:29}
\tau = \rho^{-1}\psi\rho
\end{equation}
\\
equation (\ref{eq:TPF3}) becomes
\begin{equation}
\label{eq:TPF4}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi\in S_{2q}}\mathrm{Tr}\big(\hat{P}_{[S]}\Gamma_{R}(\rho^{-1}\psi\rho)\big)C^{4\mu}_{K}C^{L}_{4\mu}(\psi)^{K}_{L}\\
\end{equation}
\\
With this relabelling, we are now contracting indices $1\&2$, $3\&4$ $\cdots (2q-1\&2q)$,
\begin{equation}
\label{eq:30}
C^{4\mu}_{K}C^{L}_{4\mu}(\psi)^{K}_{L} = (\psi)^{k_{1}k_{1}k_{2}k_{2}\cdots k_{q}k_{q}}_{l_{1}l_{1}l_{2}l_{2}\cdots l_{q}l_{q}}
\end{equation}
\\
Inside the trace, $\Gamma_{R}(\rho)\hat{P}_{[S]}\Gamma_{R}(\rho^{-1}) $ is now a sum over $S_{q}[S_{2}]$ which stabilises the set $(1,2),(3,4)\cdots (2q-1,2q)$. We now have the following for the two-point function
\begin{equation}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi\in S_{2q}}\mathrm{Tr}\big(P_{[S]}\Gamma_{R}(\psi)\big)C^{4\mu}_{K}C^{L}_{4\mu}(\psi)^{K}_{L}\\
\end{equation}
\\
Now split the sum over $S_{2q}$ up into a sum over the Brauer algrebra $\mathcal{B}_{q}$ and a sum over $S_{q}[S_{2}]$ as was done in \cite{SON1}. $\mathcal{B}_{q}$ is isomorphic to the coset $S_{2q}/S_{q}[S_{2}]$.
\begin{equation}
\label{eq:TPF5}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta}\frac{(2q)!n!m!2^{q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi_{1}\in \mathcal{B}_{q}}\sum\limits_{\psi_{2}\in S_{q}[S_{2}]}\mathrm{Tr}\big(P_{[S]}\Gamma_{R}(\psi_{1}\psi_{2})\big)C^{4\mu}_{K}C^{L}_{4\mu}(\psi_{1}\psi_{2})^{K}_{L}
\end{equation}
\\
Now $\psi_{2}$ is an element of the stabiliser of $\sigma_{4\mu}$, which means that
\begin{equation}
\label{eq:31}
C^{4\mu}_{K}(\psi_{2})^{K}_{I} = C^{4\mu}_{I}
\end{equation}
\\
Then
\begin{equation}
\label{eq:32}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta}\frac{(2q)!q!n!m!2^{2q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\sum\limits_{\psi_{1}\in \mathcal{B}_{q}}\mathrm{Tr}\big(\Gamma_{R}(\psi_{1})P_{[S]}\big)C^{4\mu}_{I}C^{L}_{4\mu}(\psi_{1})^{I}_{L}
\end{equation}
\\
Since $\mathcal{B}_{q}$ is the set of all coset representatives, and therefore not unique, we again choose the following elements for $\mathcal{B}_{q}$ \cite{SON1}\footnote{Indeed, this is why we changed the embedding in the first place - so we could use the permutations in (\ref{eq:CosetReps}) as the set of coset representatives.}.
\begin{equation}
\label{eq:CosetReps}
\prod^{q-1}_{j=0}\prod^{2j+1}_{i=1}(i,2j+1)
\end{equation}
\\
It then follows that
\begin{equation}
\label{eq:TPFformula}
\langle O_{R(r,s)\alpha}\overline{O}_{T(t,u)\beta} \rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!q!n!m!2^{2q}}{d_{R}((2n)!(2m)!)^{2}}\bra{[S]}\big(\prod^{q-1}_{j=0}(N+J_{2j+1})\ket{[S]}
\end{equation}
\\
where we wrote $P_{[S]} = \ket{[S]}\!\!\bra{[S]}$. $J_{2j+1}$ is the Jucys-Murphy element in irrep $R$. $J_{2j+1}$ acting on a state in $R$ returns the content of the box labeled $2j+1$. Thus, $N + j_{2j+1}$ gives the weight of that box. Keep in mind that the $\ket{[S]}$ in (\ref{eq:TPFformula}) is symmetric in boxes $(1,2)$ and $(3,4) \cdots$, whereas the $\ket{[S]}$ in the definition of the restricted Schurs is symmetric in boxes $(1,4)$ and $(2,3)$ and so on.
The state $\ket{[S]}$ in (\ref{eq:TPFformula}) only consists of states in $R$ that has boxes $1\&2$ next to each other, and $3\&4$ next to each other $\cdots$ and boxes $2q-1\&2q$ next to each other. As examples, we have
\begin{equation}
\label{eq:33}
\begin{Young} 1&2&3&4\cr5&6\cr7&8\cr \end{Young} \hspace{20pt} and \hspace{20pt}\begin{Young} 1&2&5&6\cr3&4\cr7&8\cr \end{Young}
\end{equation}
\\
contributing to $\ket{[S]}$ for this particular $R$. In the end, it is only boxes in the odd columns whose weights contribute to the product in (\ref{eq:TPFformula}). By odd columns, we mean the weights of the all the boxes in the 1st column and the weights of all the boxes in the 3rd column and then the 5th column and so on. For example we take the weights of the starred boxes
\begin{equation}
\label{eq:34}
\begin{Young} *&\cr*&\cr \end{Young},\;\;\;\begin{Young} *&&*&\cr*&&*&\cr \end{Young}, \;\;\;\begin{Young} *&\cr*&\cr*&\cr*&\cr \end{Young}, \;\;\; \begin{Young} *&&*&\cr*&\cr*&\cr \end{Young}
\end{equation}
\\
We then arrive at the following formula for the two-point function
\begin{equation}
\label{eq:Finalformula}
\langle O_{R(r,s)\alpha} \overline{O}_{T(t,u)\beta}\rangle = \delta_{RT}\delta_{rt}\delta_{su}\delta_{\alpha\beta} \frac{(2q)!q!n!m!2^{2q}}{d_{R}\big( (2n)!(2m)! \big)^{2}}\prod\limits_{i\;\in\; odd\;columns\;in\;R} c_{i}
\end{equation}
\\
We test our formula for a variety of $S_{4}$ and $S_{8}$ examples in appendix 2. Unfortunately, the first multiplicity in the $S_{2n}\times S_{2m}$ irreps we found occurred for the case of
\begin{equation}
R = \begin{Young} &&&&&\cr&&&\cr&&&\cr&\cr&\cr&\cr & \cr \end{Young} \hspace{20pt} \mathrm{subducing\;2\;copies\;of} \hspace{20pt} (r,s) = ( \begin{Young} &&&\cr&&&\cr&\cr&\cr\cr \cr \end{Young}, \begin{Young} &&\cr&&\cr\cr \cr\end{Young})
\end{equation}
\\
which is beyond our computational capability.
\section{Discussion}
In this work, we have constructed an exactly orthogonal basis for the $1/4$-BPS sector of free $\mathcal{N}=4$ super Yang-Mills theory with $SO(2n)$ gauge group. We have shown that the counting of states in this sector exactly matches the number of restricted Schurs the can be defined for $(r,s)$ having even columns and $R$ having even rows. Furthermore, the counting went according to $g(r,s,R)$. This means that there are less operators than for the $U(N)$ case for which the counting went according to $g(r,s,R)^{2}$. We have also presented a variety of simple examples in support of our findings. We then constructed a basis of operators which matched the counting. The basis we constructed also has the nice symmetry property of being invariant under sending
\begin{equation}
\label{eq:35}
\sigma \rightarrow \eta \sigma \xi \hspace{20pt} \eta \in S_{n}[S_{2}]\times S_{m}[S_{2}] \; \mathrm{and}\; \xi \in S_{q}[S_{2}].
\end{equation}
\\
Finally we achieved an analytic formula for the two-point function in the form of (\ref{eq:Finalformula}). The restricted Schurs (\ref{eq:RSDef}) are orthogonal and its two-point function is relatively simple to evaluate. We take the product of the weights of the boxes in the odd columns. We also computed a few simple, but non-trivial, examples to check this formula. A next natural step is to study the spectrum of anomalous dimensions of these operators. Using AdS/CFT, we can then study physics of the dual brane objects. We hope to make progress in this direction soon.\\
\emph{Acknowledgements:} I would like to thank Pablo Diaz, Nirina Tahiridimbisoa for many useful discussions. In particular, I'd like to thank Prof. Robert de Mello Koch for all the input and encouragement, and without whom this work would certainly not have been possible.
\section{Appendix 1. Counting examples}
\subsubsection{1 $Z$ and 1 $Y$ for $SO(4)$}
First consider the case of having only 1 $Z$ and 1 $Y$ for $SO(4)$. The partition function gives a total of 2 operators. We now try to match this number with the number of restricted Schurs that can be defined. There is only 1 possible diagram for $r$,
\begin{equation}
\label{eq:36}
\begin{Young} \cr\cr \end{Young}
\end{equation}
\\
and one possible Young diagram for $s$,
\begin{equation}
\label{eq:37}
\begin{Young} \cr\cr \end{Young}
\end{equation}
\\
Multiplying $r$ and $s$ together, we find
\begin{equation}
\label{eq:38}
\begin{Young} \cr\cr \end{Young}\otimes \begin{Young} 1\cr2\cr \end{Young} =
\begin{Young}& 1\cr&2\cr \end{Young} \;\otimes \; \begin{Young} \cr\cr1\cr2\cr \end{Young}
\end{equation}
\\
Using $\delta_{ij}$ and $\epsilon_{ijkl}$, we can construct only are two operators:
\begin{eqnarray}
\label{eq:39}
\mathcal{O}_{1} &=& \mathrm{Tr}(ZY)
\label{eq:40}\\
\mathcal{O}_{2} &=& \epsilon_{ijkl}Z^{ij}Y^{kl}
\end{eqnarray}
\\
Thus, we have the following correspondence
\begin{eqnarray}
\mathrm{Tr}(ZY) &\leftrightarrow& \begin{Young}& \cr&\cr \end{Young}, (\begin{Young}\cr\cr\end{Young},\begin{Young}\cr\cr\end{Young})\\
\epsilon_{ijkl}Z^{ij}Y^{kl} &\leftrightarrow& \begin{Young} \cr\cr\cr\cr \end{Young}, (\begin{Young}\cr\cr\end{Young},\begin{Young}\cr\cr\end{Young})
\end{eqnarray}
\\
where the last operator is known as the Pfaffian \cite{SON1}
\subsubsection{2 $Z$'s and 2 $Y$'s for $SO(4)$}
For 2 $Z$'s and 2 $Y$'s for $SO(4)$, the partition function gives a total of 7 operators. We thus expect 7 different restricted Schur labels. For $r$ we can have
\begin{equation}
\label{eq:41}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
For $s$ we can have
\begin{equation}
\label{eq:42}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying $r$ and $s$ together and only taking the results with even rows, we get
\begin{eqnarray}
\begin{Young} &\cr&\cr \end{Young} \otimes \begin{Young} 1&1\cr2&2\cr \end{Young} &\mathrm{gives}&
\begin{Young} &&1&1\cr&&2&2\cr \end{Young}\;,\; \begin{Young} &&1&1\cr&\cr2&2\cr \end{Young} \;,\; \begin{Young} &&1\cr&&2\cr1\cr2\cr \end{Young}\;
\;,\;\begin{Young} &\cr&\cr 1&1 \cr2 &2\cr \end{Young}
\end{eqnarray}
\\
Next
\begin{eqnarray}
\begin{Young} &\cr&\cr \end{Young} \otimes \begin{Young} 1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr3\cr4\cr \end{Young}
\end{eqnarray}
\\
Then
\begin{eqnarray}
\begin{Young} \cr\cr\cr\cr \end{Young} \otimes \begin{Young} 1&1\cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr \end{Young}
\end{eqnarray}
\\
Finally
\begin{eqnarray}
\begin{Young} \cr\cr\cr\cr \end{Young} \otimes \begin{Young} 1\cr2\cr3\cr4\cr \end{Young}&\mathrm{gives}& \begin{Young} &1\cr&2\cr&3 \cr &4\cr \end{Young}
\end{eqnarray}
\\
Counting all the Young diagrams we get 7 with the correct labelling. We also check that we find 7 operators using $\delta_{ij}$ and $\epsilon_{ijkl}$. They are listed below
\begin{eqnarray}
\mathcal{O}_{1}&=& \mathrm{Tr}(Z^{2}Y^{2}) \nonumber \\
\mathcal{O}_{2}&=& \mathrm{Tr}(ZYZY) \nonumber \\
\mathcal{O}_{3}&=& \mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2})\nonumber \\
\mathcal{O}_{4}&=& \mathrm{Tr}(ZY)^{2} \nonumber \\
\mathcal{O}_{5} &=& \epsilon_{ijkl}Z^{ij}Z^{kl}\mathrm{Tr}(Y^{2}) \nonumber \\
\mathcal{O}_{6} &=& \epsilon_{ijkl}Z^{ij}Y^{kl}\mathrm{Tr}(ZY) \nonumber \\
\mathcal{O}_{7} &=& \epsilon_{ijkl}Y^{ij}Y^{kl}\mathrm{Tr}(Z^{2})
\end{eqnarray}
\\
\subsubsection{3 $Z$'s and 2 $Y$'s for $SO(6)$}
Here we do not expect multitrace operators. This is because there is an odd number of fields; we will always end up with a $\mathrm{Tr}(Z)$ or $\mathrm{Tr}(Y)$ which vanishes. We can, however, obtain operators built using the $\epsilon_{ijklmn}$. The partition function gives a total of 4 operators. We thus expect 4 restricted Schurs and each of the should have the Pfaffian as a factor. For $r$, the possible Young diagrams are
\begin{equation}
\label{eq:43}
\begin{Young} &&\cr&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
and for $s$, the possible Young diagrams are
\begin{equation}
\label{eq:44}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying each $r$ with each $s$, and taking only the diagrams which contribute, we find
\begin{eqnarray}
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& 0 \\
&& \nonumber \\
\begin{Young} &\cr&\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr\cr\cr1\cr2\cr \end{Young} \\
&& \nonumber \\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr\cr\cr \end{Young}\\
&&\nonumber \\
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&\cr&&\cr1\cr 2\cr3\cr4\cr \end{Young}\\
&& \nonumber \\
\begin{Young} &\cr&\cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr\cr\cr3\cr4\cr \end{Young}
\end{eqnarray}
\\
\begin{eqnarray}
&& \nonumber \\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}\otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0
\end{eqnarray}
\\
We indeed obtain 4 restricted Schurs as expected.
\subsubsection{4 $Z$'s and $2$Y's for SO(8)}
The partition function gives a total of 13 operators. The possible diagrams for $r$ are
\begin{equation}
\label{eq:45}
\begin{Young} &&&\cr&&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr&\cr&\cr \end{Young} , \hspace{20pt} \begin{Young} &&\cr&&\cr\cr \cr \end{Young} , \hspace{20pt} \begin{Young} &\cr&\cr\cr \cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
and for $s$
\begin{equation}
\label{eq:46}
\begin{Young}& \cr&\cr \end{Young}, \hspace{20pt} \begin{Young}\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying each $r$ with each $s$ and taking only the diagrams which contribute to the counting, we find
\begin{eqnarray}
\begin{Young} &&&\cr&&&\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&&&1&1\cr&&&&2&2\cr \end{Young} \;,\; \begin{Young} &&&&1&1\cr&&&\cr2&2\cr \end{Young}\;,\;\begin{Young} &&&\cr&&&\cr 1&1\cr2&2\cr\end{Young} \nonumber \\
\begin{Young} &&&\cr&&&\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0 \nonumber \\
\begin{Young} &\cr&\cr&\cr&\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1&1\cr&&2&2\cr&\cr&\cr \end{Young}\;,\; \begin{Young} &&1&1\cr&\cr&\cr&\cr2&2\cr \end{Young}\;,\;\begin{Young} &\cr&\cr&\cr&\cr 1&1\cr2 &2\cr\end{Young} \nonumber \\
\begin{Young} &\cr&\cr&\cr&\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0\\
\begin{Young}& &\cr&&\cr\cr \cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young}&&&1\cr&&&2\cr&1\cr&2 \cr \end{Young} \nonumber \\
\begin{Young}&& \cr&& \cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young}&&&1 \cr&&&2 \cr&3\cr&4\cr \end{Young} \;,\; \begin{Young}&& \cr&& \cr\cr\cr1\cr2\cr3\cr4 \cr\end{Young} \nonumber \\
\begin{Young}& \cr& \cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young}&&1 \cr&&2 \cr\cr\cr\cr\cr1\cr2\cr \end{Young}\nonumber
\end{eqnarray}
\\
\begin{eqnarray*}
\begin{Young}& \cr& \cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}&
\begin{Young}&&1 \cr&&2 \cr\cr\cr\cr\cr3\cr4\cr \end{Young}\;,\;\begin{Young}& \cr& \cr&1\cr&2\cr&3\cr&4\cr \end{Young} \\
\begin{Young} \cr\cr\cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1&1 \cr2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr\cr\cr\cr\cr \end{Young}\\
\begin{Young} \cr\cr\cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young}1\cr2\cr3\cr4\cr \end{Young} &\mathrm{gives}& 0
\end{eqnarray*}
\\
Adding up all the Young diagram labels, we indeed get 13 operators. The reader is invited to check that for $SO(4)$, there are 12 operators and for $SO(6)$, there are 9 operators.
\subsubsection{3 $Z$'s and 3 $Y$'s for $SO(8)$}
As a final example, we consider 3 $Z$'s and 3 $Y$'s for $SO(8)$. The partition function gives a total of 14 operators. For $r$, we have
\begin{equation}
\label{eq:47}
\begin{Young} &&\cr&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
and for $s$
\begin{equation}
\label{eq:48}
\begin{Young} &&\cr&&\cr \end{Young}, \hspace{20pt} \begin{Young} &\cr&\cr\cr\cr \end{Young}, \hspace{20pt} \begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}
\end{equation}
\\
Multiplying, we find
\begin{eqnarray}
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young} 1&1&1\cr2&2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&&1&1&1\cr&&&2&2&2\cr \end{Young} \;,\; \begin{Young} &&&1&1&1\cr&&&2\cr2&2\cr \end{Young},\; \begin{Young} &&&1\cr&&&2\cr1&1\cr2&2\cr \end{Young} \nonumber \\
\begin{Young} &\cr&\cr\cr\cr \end{Young}\otimes \begin{Young} 1&1&1\cr2&2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1&1\cr&&2&2\cr&1\cr&2\cr \end{Young} \nonumber \\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young} \otimes \begin{Young} 1&1&1\cr2&2&2\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1 \cr&2&2\cr\cr\cr\cr\cr1\cr2\cr \end{Young} \nonumber \\
\begin{Young} &&\cr&&\cr \end{Young} \otimes \begin{Young} 1&1\cr2&2\cr.\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &&&1\cr&&&2\cr1&3\cr2&4\cr \end{Young}\\
\begin{Young} &\cr&\cr\cr\cr \end{Young}\otimes \begin{Young} 1&1\cr2&2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young}&&1&1\cr&&2&2\cr&3\cr&4\cr \end{Young}\;,\; \begin{Young} &&1&1\cr&\cr&2\cr&3\cr2&4\cr \end{Young}\;,\; \begin{Young} &&1\cr&&2\cr\cr\cr1\cr2\cr3\cr4\cr \end{Young} \;,\; \begin{Young} &\cr&\cr&1\cr&2\cr1&3\cr2&4\cr \end{Young} \nonumber
\end{eqnarray}
\\
\begin{eqnarray*}
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}\otimes \begin{Young} 1&1\cr2&2\cr3\cr4\cr \end{Young} &\mathrm{gives}& \begin{Young} &1&1\cr&2&2\cr\cr\cr\cr\cr3\cr4\cr \end{Young}\\
\begin{Young} &&\cr&&\cr \end{Young}\otimes \begin{Young} 1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} &\mathrm{gives}& \begin{Young} &&\cr&&\cr1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} \\
\begin{Young} &\cr&\cr\cr\cr \end{Young}\otimes \begin{Young} 1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} &\mathrm{gives}& \begin{Young} &&1\cr&&2\cr\cr\cr3\cr4\cr5\cr6\cr \end{Young}\\
\begin{Young} \cr\cr\cr\cr\cr\cr \end{Young}\otimes \begin{Young} 1\cr2\cr3\cr4\cr5\cr6\cr \end{Young} &\mathrm{gives}& \begin{Young} &1\cr&2\cr&3\cr&4\cr&5\cr&6\cr \end{Young}
\end{eqnarray*}
\\
Adding up all the Young diagram labels, we get our expected 14 restricted Schur polynomials. Again the reader is invited to check that for $SO(6)$ there are 10 restricted Schurs and for $SO(4)$, there are 12 restricted Schurs.
\section{Appendix 2. Two-point function examples}
In the following examples, we denote the eigenvector of $\hat{P}_{[S]}$ by $\hat{\ket{[S]}}$.
\subsection{$q=2$}
For $q=2$, we have
\begin{equation}
\label{eq:50}
R = \begin{Young}&\cr&\cr\end{Young}, \hspace{20pt} (r,s) = (\begin{Young}\cr\cr\end{Young},\begin{Young}\cr\cr\end{Young})
\end{equation}
\\
Firstly
\begin{equation}
\label{eq:51}
\ket{[\hat{S}]} = \frac{1}{2}\,\begin{Young}4&3\cr2&1\cr\end{Young} + \frac{\sqrt{3}}{2}\,\begin{Young}4&2\cr3&1\cr\end{Young}, \hspace{20pt} \ket{[A],[A]} = \begin{Young}4&2\cr3&1\cr\end{Young}
\end{equation}
\\
The restricted Schur is
\begin{equation}
\label{eq:52}
O_{R(r,s)} = 2\sqrt{3}\mathrm{Tr}(ZY)
\end{equation}
\\
Evaluating the Wick contractions, its two point function is
\begin{equation}
\label{eq:53}
\langle O_{R(r,s)}\overline{O}_{T(t,u)} \rangle = 12 \langle \mathrm{Tr}(ZY) \mathrm{Tr}(\overline{Z}\overline{Y}) \rangle = 24N(N-1)
\end{equation}
\\
Using (\ref{eq:TPFformula}), with $q=2, n=m=1$ and $d_{R}=2$, we get
\begin{equation}
\label{eq:54}
\langle O_{R(r,s)}\overline{O}_{T(t,u)} \rangle = 24N(N-1).
\end{equation}
\\
\subsection{$q=4$}
First consider
\begin{equation}
\label{eq:55}
R = \begin{Young} &&&\cr&&&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &\cr&\cr \end{Young},\begin{Young} &\cr&\cr \end{Young})
\end{equation}
\\
We caluculate
\begin{equation}
\label{eq:ShatforR44}
\ket{[\hat{S}]} = \frac{\sqrt{5}}{4} \begin{Young} 8&6&4&2\cr7&5&3&1\cr \end{Young} + \sqrt{\frac{5}{48}} \begin{Young} 8&7&4&2\cr6&5&3&1\cr \end{Young} + \frac{\sqrt{5}}{12} \begin{Young} 8&7&4&3\cr6&5&2&1\cr \end{Young} + \frac{2}{3} \begin{Young} 8&7&6&5\cr4&3&2&1\cr \end{Young} + \sqrt{\frac{5}{48}}\begin{Young} 8&6&4&3\cr7&5&2&1\cr \end{Young}
\end{equation}
\\
and
\begin{equation}
\label{eq:AAvecforR44}
\ket{[A],[A]} = \begin{Young} 8&6&4&2\cr7&5&3&1\cr \end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:56}
O_{R(r,s)} = \frac{2\sqrt{5}}{3}\Big( 2\mathrm{Tr}(ZY)^{2} + 2\mathrm{Tr}(ZYZY) + \mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2}) + 4\mathrm{Tr}(Z^{2}Y^{2}) \Big)
\end{equation}
\\
and, after evaluating all the Wick contractions, its two-point function is
\begin{equation}
\label{eq:57}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N+1)(N+2)
\end{equation}
\\
Equation (\ref{eq:TPFformula}), for $q=4, n=m=2, d_{R} = 14$ and taking only the weights of the odd columns, gives
\begin{equation}
\label{eq:58}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N+1)(N+2)
\end{equation}
\\
Next consider
\begin{equation}
\label{eq:59}
R = \begin{Young} &\cr&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &\cr&\cr \end{Young},\begin{Young} &\cr&\cr \end{Young})
\end{equation}
\\
The state $\ket{[\hat{S}]}$ was found to be
\begin{equation}
\label{eq:60}
\ket{[\hat{S}]} = \frac{3}{4} \begin{Young}8 &6\cr7&5\cr4&2\cr3&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young}8 &7\cr6&5\cr4&2\cr3&1\cr \end{Young} + \frac{1}{4} \begin{Young}8 &7\cr6&5\cr4&3\cr2&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young} 8&6\cr7&5\cr4&3\cr2&1\cr \end{Young}
\end{equation}
\\
and
\begin{equation}
\label{eq:61}
\ket{[A],[A]} = \begin{Young}8 &6\cr7&5\cr4&2\cr3&1\cr \end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:62}
O_{R(r,s)} = \frac{2}{3}\Big( 2\mathrm{Tr}(ZY)^{2} + 2\mathrm{Tr}(ZYZY) + 3\mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2}) - 12\mathrm{Tr}(Z^{2}Y^{2}) \Big)
\end{equation}
\\
and its two-point function is
\begin{equation}
\label{eq:63}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
Equation (\ref{eq:TPFformula}), for $q=4, n=m=2, d_{R} = 14$ and taking the weights from the odd columns, gives
\begin{equation}
\label{eq:64}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
Next consider
\begin{equation}
\label{eq:65}
R = \begin{Young} &\cr&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} \cr\cr\cr\cr \end{Young},\begin{Young} \cr\cr\cr\cr \end{Young})
\end{equation}
\\
We find
\begin{equation}
\label{eq:ShatforR2222}
\ket{[\hat{S}]} = \frac{3}{4} \begin{Young}8 &6\cr7&5\cr4&2\cr3&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young}8 &7\cr6&5\cr4&2\cr3&1\cr \end{Young} + \frac{1}{4} \begin{Young}8 &7\cr6&5\cr4&3\cr2&1\cr \end{Young} + \frac{\sqrt{3}}{4} \begin{Young} 8&6\cr7&5\cr4&3\cr2&1\cr \end{Young}
\end{equation}
\\
and
\begin{equation}
\label{eq:66}
\ket{[A],[A]} = \begin{Young}8 &4\cr7&3\cr6&2\cr5&1\cr \end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:67}
O_{R(r,s)} = \frac{4\sqrt{5}}{3}\Big( \mathrm{Tr}(ZY)^{2} - 2\mathrm{Tr}(ZYZY) \Big)
\end{equation}
\\
and its two-point function is
\begin{equation}
\label{eq:68}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
Equation (\ref{eq:TPFformula}) gives
\begin{equation}
\label{eq:69}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{640}{3}N(N-1)(N-2)(N-3)
\end{equation}
\\
For the final example for which $n=m=2$, consider
\begin{equation}
\label{eq:6911}
R = \begin{Young} &&&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &\cr&\cr \end{Young},\begin{Young} &\cr&\cr \end{Young})
\end{equation}
\\
We find the following for $\ket{[\hat{S}]}$ and $\ket{[A],[A]}$
\begin{eqnarray}
\ket{[\hat{S}]} &=& \frac{1}{4\sqrt{6}}\begin{Young} 8&7&2&1\cr6&5\cr4&3\cr \end{Young} + \frac{1}{4\sqrt{2}}\begin{Young} 8&6&2&1\cr7&5\cr4&3\cr \end{Young} + \frac{1}{8\sqrt{3}}\begin{Young} 8&7&3&1\cr6&5\cr4&2\cr \end{Young} + \frac{1}{8}\begin{Young} 8&6&3&1\cr7&5\cr4&2\cr \end{Young} + \frac{\sqrt{5}}{8}\begin{Young} 8&7&4&1\cr6&5\cr3&2\cr \end{Young} \nonumber \\
&&+\; \frac{\sqrt{15}}{8}\begin{Young} 8&6&4&1\cr7&5\cr3&2\cr \end{Young} - \frac{1}{12\sqrt{2}}\begin{Young} 8&7&4&3\cr6&5\cr2&1\cr \end{Young} - \frac{1}{4\sqrt{6}}\begin{Young} 8&6&4&3\cr7&5\cr2&1\cr \end{Young} - \frac{1}{8\sqrt{3}}\begin{Young} 8&7&4&2\cr6&5\cr3&1\cr \end{Young} - \frac{1}{8}\begin{Young} 8&6&4&2\cr7&5\cr3&1\cr \end{Young} \nonumber \\
&&+\;\frac{\sqrt{5}}{8}\begin{Young} 8&7&3&2\cr6&5\cr4&1\cr \end{Young} + \frac{\sqrt{15}}{8}\begin{Young} 8&6&3&2\cr7&5\cr4&1\cr \end{Young} + \frac{1}{6}\sqrt{\frac{5}{2}}\begin{Young} 8&7&6&5\cr4&3\cr2&1\cr \end{Young} + \frac{1}{2}\sqrt{\frac{5}{6}}\begin{Young} 8&7&6&5\cr4&2\cr3&1\cr \end{Young}
\end{eqnarray}
\\
and
\begin{equation}
\label{eq:70}
\ket{[A],[A]} = -\frac{3}{8}\begin{Young} 8&6&3&1\cr7&5\cr4&2\cr \end{Young} + \frac{\sqrt{15}}{8}\begin{Young} 8&6&4&1\cr7&5\cr3&2\cr \end{Young} - \frac{5}{8}\begin{Young} 8&6&4&2\cr7&5\cr3&1\cr \end{Young} + \frac{\sqrt{15}}{8}\begin{Young} 8&6&3&2\cr7&5\cr4&1\cr\end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:71}
O_{R(r,s)} = \frac{4}{3}\Big( -\mathrm{Tr}(ZY)^{2} - \mathrm{Tr}(ZYZY) + \mathrm{Tr}(Z^{2})\mathrm{Tr}(Y^{2}) + \mathrm{Tr}(Z^{2}Y^{2}) \Big)
\end{equation}
\\
and its two-point function, after evaluating all the Wick contractions, is
\begin{equation}
\label{eq:72}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{160}{3}N(N-1)(N-2)(N+2)
\end{equation}
\\
Equation (\ref{eq:TPFformula}), for $q=4, n=m=2, d_{R} = 56$ and odd columns, gives
\begin{equation}
\label{eq:73}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{160}{3}N(N-1)(N-2)(N+2)
\end{equation}
\\
Now let's try two examples where $n=3, m=1$, i.e., the number of $Z$'s is not the equal to the number of $Y$'s. Consider
\begin{equation}
\label{eq:74}
R = \begin{Young} &&&\cr&&&\cr \end{Young}, \hspace{20pt} (r,s) = (\begin{Young} &&\cr&&\cr \end{Young},\begin{Young} \cr\cr \end{Young})
\end{equation}
\\
States $\hat{\ket{[S]}}$ and $\ket{[A],[A]}$ are still the same as in (\ref{eq:ShatforR44}) and (\ref{eq:AAvecforR44}). The operator is
\begin{equation}
\label{eq:75}
O_{R(r,s)} = \frac{4}{\sqrt{5}}\Big( \mathrm{Tr}(ZY)\mathrm{Tr}(Z^{2}) + 2 \mathrm{Tr}(Z^{3}Y) \Big)
\end{equation}
\\
and its two-point function is
\begin{equation}
\label{eq:76}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{256}{5}N(N-1)(N+1)(N+2).
\end{equation}
\\
Formula (\ref{eq:TPFformula}) reproduces exactly this result. As the final example, we consider
\begin{equation}
\label{eq:77}
R = \begin{Young} &\cr&\cr&\cr&\cr \end{Young}, \hspace{20pt} (r,s) = ( \begin{Young} &\cr&\cr\cr\cr \end{Young}, \begin{Young} \cr\cr \end{Young})
\end{equation}
\\
The $\hat{\ket{[S]}}$ is the same as in (\ref{eq:ShatforR2222}). However
\begin{equation}
\label{eq:78}
\ket{[A],[A]} = \frac{\sqrt{5}}{3}\begin{Young} 8&4\cr7&3\cr6&2\cr5&1\cr\end{Young} + \frac{2}{3}\begin{Young} 8&6\cr7&5\cr4&2\cr3&1\cr\end{Young}
\end{equation}
\\
The operator is
\begin{equation}
\label{eq:79}
O_{R(r,s)} = \frac{4}{\sqrt{5}}\Big( \mathrm{Tr}(ZY)\mathrm{Tr}(Z^{2}) - 2 \mathrm{Tr}(Z^{3}Y) \Big)
\end{equation}
\\
The two-point function is
\begin{equation}
\label{eq:80}
\langle O_{R(r,s)}\overline{O}_{R(r,s)} \rangle = \frac{256}{5}N(N-1)(N-2)(N-3).
\end{equation}
\\
which, again, is exactly reproduced by (\ref{eq:TPFformula}).
\bibliographystyle{unsrt}
|
\section{Introduction}
\vspace{.5cm}
QCD describes interactions as the exchange of a quantum degree of freedom called ``color". The theory of strong interactions has the interesting property of asymptotic freedom. As the running coupling constant $\alpha_s(Q^2)$ becomes smaller, the perturbative treatment of QCD allows for the explanation of the hadronic phenomena, whose basic ingredients are the Parton Distributions.
On the other hand, hadrons that are actually observed in nature carry no color --- they are color singlets. So far, physicists have failed to describe colorless hadrons within QCD due to the property of confinement:
At low energy, there is no justification for a perturbative treatment of QCD. We do not longer have a description of the relevant low energy observables from QCD. In this regime, non-perturbative approaches come into play.
In these proceedings, we will describe how, in these schemes, parameters are fixed by phenomenology. The two regimes of the strong interactions form an undivided whole, though the transition of the relevant degrees of freedom is not fully understood yet. As such, low energy parameters ---as accounting for the physical phenomena--- will guide high-energy observables. Among the many interesting implications of non-perturbative physics for perturbative QCD and high-energy observables, we will consider 2 main directions. The first is related to the PDFs. The quark degrees of freedom transition into hadronic degrees of freedom at a scale intimately related to the initial scale chosen for the PDF fits. We here discuss various approaches of the determination of the transition scale in Sections~\ref{sect:hadro_scale} $\&$ \ref{sect:NP_scale}. The first steps toward ``soft" evolution are discussed in Section~\ref{sect:NP_scale}. The phenomenological role of the running coupling constant in the infrared region is highlighted in Section~\ref{sect:soft_evo}.
This analysis induced a need to reconsider the PDFs at large values of Bjorken-$x$, close to the exclusive limit. Interestingly, the large-$x$ PDF carry important consequences at the high-energy level as is shown in Section~\ref{sect:largex}.
The second direction we want to explore, in Section~\ref{sect:charges}, relates hadronic matrix elements to observables of physics Beyond the Standard Model. In particular, the scalar and tensor charges have been shown to be of great interest for precision measurements of new physics.
\section{Parton Distribution Functions at High Energy}
\vspace{.5cm}
A way of connecting the perturbative and non-perturbative worlds has traditionally been through the study of Parton Distribution Functions (PDFs):
Deep Inelastic processes are such that they enable us to look with a good resolution inside the hadron and allow us to resolve the very short distances, {\it i.e.} small configurations of quarks and gluons. This part of the process is described through perturbative QCD. A resolution of such short distances is obtained with the help of non-strongly interacting probes. Such a probe, typically a photon, is provided by hard reactions. In that scheme, the PDFs reflect how the target reacts to the probe, or how the quarks and gluons are distributed inside the target. The insight into the structure of hadrons is reached at that stage: the large virtuality of the photon, $Q^2$, involved in such processes allows for the factorization of the hard (perturbative) and soft (non-perturbative) contributions in their amplitudes ---in an Operator Product Expansion style. They are matrix elements of a bilocal current on the light-cone,
\begin{eqnarray}
P^{\mu} q(x)&=& \int \frac{d\tau}{4\pi}\, e^{ix\tau} \left \langle PS\right | {\bar q}(0) \gamma^{\mu} q(\tau n)\left |PS\right\rangle\quad,
\label{eq:pdf}
\end{eqnarray}
with $n^{\mu}=(1,0,0,-1)/(\sqrt{2} \Lambda)$ a light-like vector. The distribution functions are non-perturbative objects as they describe the large distance behavior of hadrons, a regime where confinement starts to matter.
The virtuality of the photon introduces {\it the factorization scale}, {\it i.e.} PDFs explicitly depend on $Q^2$.
This $Q^2$-evolution is dictated by the Dokshitzer-Gribov-Lipatov-Altarelli-Parisi (DGLAP) equations,
%
\begin{eqnarray}
Q^2\,\frac{\partial}{\partial Q^2} q(x, Q^2)&=& \frac{\alpha_s(Q^2)}{2\pi}\, \int_x^1 \frac{d\xi}{\xi}\, P\left(\frac{x}{\xi}, \alpha_s(Q^2)\right)\, q(\xi, Q^2)\quad,
\label{eq:DGLAP}
\end{eqnarray}
here at leading order in $\alpha_s$--- and where $P(x, Q^2)$ are the splitting functions.
At leading order ({\it leading-twist}), there are three types of
PDFs: the ordinary number density, the helicity, and the transversity. The former, $q(x)$, is the well-known unpolarized PDF and is accessible through, {\it e.g.}, inclusive deep inelastic scattering (DIS), {\it i.e.} in the Bjorken regime.
On the other hand, $g_1(x)$, the helicity PDF, is less constrained. The experimental knowledge on $h_1(x)$, the transversity PDF, is sparse as it is a chiral-odd quantity, not accessible through fully inclusive processes.
Collinear PDFs are universal and come into play in many processes, {\it e.g.} Drell-Yan processes, proton-proton collisions, $\cdots$ and are therefore of the utmost importance, especially for precision measurements.
Unpolarized PDFs have been extensively studied from first principles, in models and fits. The global fits combine various data set with different energy range. The precision obtained in the various fits, for light flavor and valence quarks, is high ; gluon and sea quarks as well as low and large-$x$ regions can incontestably be improved. For example, in DIS,
the hard scattering part of the process
can be presently described using splitting functions up to next-to-next-to-leading order (NNLO) \cite{Moch:2004pa,Vogt:2004mw}, and Wilson coefficients functions up to N$^3$LO
\cite{Vermaseren:2005qc}.
PDF parametrizations along with their uncertainties have been obtained applying this framework up to NNLO, by a number of collaborations (see review in Ref.~\cite{Forte:2013wc}). We cite MSTW, CTEQ, NNPDF, HERAFitter, GJR, ABM.\footnote{A useful link to compare PDF sets can be found in Ref.~\cite{durham} or the proceedings of the PDF4LHC working groups.}
The parameterizations present statistical error on the PDF itself as well as on the value of $\alpha_s(M_Z^2)$. The propagation of these uncertainties for the LHC phenomenology has been extensively studied, see {\it e.g.} \cite{Forte:2013wc}. A representative example is the Higgs production through gluon-gluon fusion: $\sim 7\%$ of uncertainty on the cross section comes from uncertainty on PDF and $\alpha_s(M_Z^2)$. Those numbers vary from one parameterization to the other.\footnote{Fig.~7 of Ref.~\cite{Forte:2013wc} is very illustrative.}
In the standard PDF fitting approaches, an {\it arbitrary} initial scale for the evolution equations, $Q_0^2>1$ GeV$^2$, is chosen: In parton distribution analyses the $x$-dependence of the PDFs is thus extracted at a particular scale $Q_0^2$, usually referred to as the {\it input scale}. A first systematic study of the effects of the choice of the input scale in global determinations of parton distributions and QCD parameters has been presented in Ref.~\cite{JimenezDelgado:2012zx}, introducing the concept of {\it procedural bias} in PDF analyses.
The first consequence of the choice of the input scale in a parton distribution analysis is that it affects the determination the strong coupling constant $\alpha_s(M_Z^2)$. In fact, the latter is determined together with the parton distributions and is substantially correlated with the gluon distribution which drives the QCD evolution.
Besides, at low scales, there is an ambiguity related to the hadronic representation. The evolution equations modify the PDFs' radiative behavior in opposition to the low-energy valence behavior. The author of Ref.~\cite{JimenezDelgado:2012zx} uses a dynamical GJR parton distribution that optimally determines the input scale~\cite{Gluck:2007ck}. The latter is used as a guideline for the corresponding degrees of freedom ---for $Q_0^2<1$ GeV$^2$ the PDF should tend to a valence-like behavior.
$Q_0^2$ turns out to be of the order of $0.55$GeV$^2$ (with $\Lambda_{\mbox{\tiny NLO}}^{n_f=3}\sim303$MeV), {\it i.e.}, in the region where non-perturbative inputs cannot be neglected, in a top-down approach.
In the next Sections, we will see how this non-perturbative input can come into play.
\section{Hadronic Scale}
\label{sect:hadro_scale}
\vspace{.5cm}
It is still a challenge to describe consistently the dynamics of scattering processes and hadronic structure at moderate energy scales. Because at such a moderate scale the hadronic representation gives way to the partonic description, it is called {\it the hadronic scale}. The hadronic scale is peculiar to each hadronic representation and should be related to the {\it input scale} defined in the previous Section.
\subsection{Determination of $Q_0$: Standard Approach}
\vspace{.5cm}
From a bottom-up point of view, the evaluation of PDFs is guided by a standard scheme, set up in valuable litterature of the 90s \cite{Traini:1997jz,Stratmann:1993aw,Parisi:1976fz}. This scheme runs in 3 main steps.
First, we either build models consistent with QCD in a moderate energy range, typically the hadronic scale; or we use effective theories of QCD for the description of hadrons at the same energy range. Second, PDFs are evaluated in these models, giving a description of the Bjorken-$x$ dependence of the distribution. Third, the scale dependence of these distributions is studied. The last step allows to bring the moderate energy description of hadrons to the {\it factorization scale}, thanks to the QCD evolution equations~(\ref{eq:DGLAP}). Here we are interested in the matching of non-perturbative models to perturbative QCD, using experimental data.
The hadronic scale is defined at a point where the partonic content of the model, defined through the second moment of the parton distribution, is known. For instance, the CTEQ parameterization gives~\footnote{MSTW gives a similar result.}
\begin{equation}
\left \langle (u_v+d_v)(Q^2=10 \mbox{GeV}^2)) \right\rangle_{n=2}=0.36\quad,
\end{equation}
with $q_v$ the valence quark distributions and with $\langle q_v (Q^2)\rangle_n=\int_0^1 dx \, x^{n-1} \, q_v(x, Q^2)$.
Scenarios for the hadronic representation have to be chosen. In an extreme scenario, {\it i.e.}, when we assume that the partons are pure valence quarks, the second Mellin moment is evolved downward until
%
\begin{equation}
\left \langle (u_v+d_v)(\mu_0^2) \right\rangle_{n=2}=1\quad.
\end{equation}
%
The hadronic scale is found to be $\mu_0^2 \sim 0.1$ GeV$^2$.
This standard procedure to fix the hadronic (non-perturbative) scale pushes perturbative QCD to its limit.
In effect, the hadronic scale turns out to be of a few hundred MeV$^2$, where the strong coupling constant has already started approaching its Landau pole. As it will be shown hereafter, the N$^m$LO evolution converges very fast, what justifies the perturbative approach.
Consequently, the behaviour of the strong coupling constant plays a central role in the QCD evolution of parton densities. We here extend the standard procedure with the non-perturbative generalization of the QCD running coupling.
\subsection{$Q_0$ from Non-Perturbative Physics}
\vspace{.5cm}
We call perturbative evolution the renormalization group equations (RGE).
The running of the coupling constant is driven by the RGE. In QCD, $\alpha_s$ is defined by renormalization conditions imposed at a large momentum scale where the coupling is small. The running coupling constant is dimensionless, but through dimensional transmutation,
the strength of the interaction may be described by a dimensionful parameter. QCD scale, $\Lambda_{\mbox{\tiny QCD}}$, is then defined as the energy scale where the interaction strength reaches the value $1$.
At N$^m$LO the scale dependence of the coupling constant is given by
$$\frac{d \, a (Q^2)}{ d(\ln \;Q^2)} = \beta_{\mbox{\tiny N}^m\mbox{\tiny LO}}(\alpha_s) =\stackrel{m}{\sum_{k=0}} a^{k+2} \beta_k,$$
where
$a = {\alpha_s / 4 \pi}.$
We show here the solution to $k=2$, {\it i.e.}, NLO.\footnote{
$
\beta_0 = 11 - \frac{2}{3}\, n_f \quad,
\beta_1 = 102 -\frac{38}{3} \,n_f $,
where $n_f$ stands for the number of effectively massless quark flavors and $\beta_k$ denote the coefficients of the usual four-dimensional $\overline{MS}$ beta function of QCD.}
The evolution equations for the coupling constant can be integrated out exactly leading to
\begin{alignat}{2}
& \ln (Q^2/\Lambda_{\mbox{\tiny LO}}^2)&&=\frac{1}{\beta_0 a_{\mbox{\tiny LO}} } \quad ,\nonumber\\
& \ln (Q^2/\Lambda_{\mbox{\tiny NLO}}^2) &&= \; \frac{1}{\beta_0 a_{\mbox{\tiny NLO}}} + \frac{b _1}{\beta_0} \ln (\beta_0 a_{\mbox{\tiny NLO}})- \frac{b _1}{\beta_0} \ln (1 + b_1 a_{\mbox{\tiny NLO}}) \quad
\label{exact}
\end{alignat}
%
where $b_k = {\beta_k / \beta_0}$. These equations, except the first, do not admit closed form solution for the coupling constant, and we have solved them numerically. We show their solution, for the same value of $\Lambda = 250$ MeV, in Fig. \ref{aperb}.
\begin{figure}
\begin{center}
\includegraphics[scale= .6]{fig/alpha_nnlo.eps}
\includegraphics[scale= .6]{fig/alpha_np.eps}
\caption{ The running of the coupling constant. {\it Left panel}: The short dashed curve corresponds to the LO solution, the medium dotted-dashed curve to NLO solution and the tiny dashed curve to the NNLO solution ($\Lambda = 250$ MeV). {\it Right panel}: The running of the effective coupling. The dotted and dashed curves represent the non-perturbative evolution with the parameters $m_0=0.3$GeV and, respectively, for the medium dashed blue curve $\rho=1.53$, for the short dashed purple curve $\rho=2.2$. The solid curve shows the NLO evolution with $\Lambda = 250$ MeV.}
\label{aperb}
\label{comparisons}
\end{center}
\end{figure}
We see in Fig.~\ref{aperb} (left panel) that the NLO and NNLO solutions agree quite well even at very low values of $Q^2$. They agree even better if we change the value of $\Lambda$ for the NNLO slightly, confirming the fast convergence of the expansion. This analysis concludes, that even close to the Landau pole, the convergence of the perturbative expansion is quite rapid, specially if we use a different value of $\Lambda$ to describe the different orders, a feature which comes out from the fitting procedures.
{\it This fast convergence ensures that perturbative evolution can still be used at rather low scales.} However, when entering the non-perturbative regime, other mechanisms take place that influence the QCD evolution. That is what we will call here non-perturbative evolution.
It is well established by now that the QCD running coupling (effective charge)
freezes in the deep infrared.
This non-perturbative property can be understood from various non-perturbative approaches~\cite{Cornwall:1981zr,Fischer:2003rp,Aguilar:2009nf,Shirkov:1997wi,Mattingly:1992ud}, {\it e.g.} from the point of view of
the dynamical gluon mass generation \cite{Cornwall:1981zr}.\footnote{Even though the gluon is massless at the level of the fundamental QCD Lagrangian, and remains
massless to all order in perturbation theory, the non-perturbative QCD
dynamics generate an effective, momentum-dependent mass, without
affecting the local $SU(3)_c$ invariance, which remains
intact.}
At the level of the Schwinger-Dyson equations
the generation of such a mass is associated with
the existence of
infrared finite solutions for the gluon propagator,
i.e. solutions with $\Delta^{-1}(0) > 0$.
Such solutions may
be fitted by ``massive'' propagators of the form
$\Delta^{-1}(Q^2) = Q^2 + m^2(Q^2)$;
$m^2(Q^2)$ is not ``hard'', but depends non-trivially on the momentum transfer $Q^2$.
One physically motivated possibility, which we shall use in here, is the so called logarithmic mass running, which is defined by
\begin{equation}
m^2 (Q^2)= m^2_0\left[\ln\left(\frac{Q^2 + \rho m_0^2}{\Lambda^2}\right)
\bigg/\ln\left(\frac{\rho m_0^2}{\Lambda^2}\right)\right]^{-1 -\gamma}.
\label{eq:mass}
\end{equation}
Note that when $Q^2\to 0$ one has $m^2(0)=m^2_0$. Even though in principle we do not have
any theoretical constraint that would put an upper bound to the value of $m_0$,
phenomenological estimates place it in the range $m_0 \sim \Lambda - 2 \Lambda$~\cite{Bernard:1981pg,Parisi:1980jy}. The other parameters were fixed at $\rho \sim 1-4$, ${\gamma} = 1/11$ \cite{Cornwall:1981zr,Aguilar:2007ie,Aguilar:2009nf}.
The non-perturbative generalization of $\alpha_s(Q^2)$
the QCD running coupling, comes in the form
\begin{equation}
a_{\mbox{\tiny NP}}(Q^2) = \left[\beta_0 \ln \left(\frac{Q^2 +\rho m^2(Q^2)}{\Lambda^2}\right)\right]^{-1} ,
\label{alphalog}
\end{equation}
where we use the same notation as before and NP stands for Non-Perturbative. Note that its zero gluon mass limit leads to the LO perturbative
coupling constant momentum dependence.
The $m^2(Q^2)$ in the argument of the logarithm
tames the Landau pole, and $a(Q^2)$ freezes
at a finite value in the IR, namely
\mbox{$a^{-1}(0)= \beta_0 \ln (\rho m^2(0)/\Lambda^2)$} \cite{Cornwall:1981zr,Aguilar:2006gr,Binosi:2009qm} as can be seen on the right panel of Fig. \ref{aperb}.
As shown in Fig.~\ref{alpha}, the coupling constant in the perturbative and non-perturbative approaches are close in size for reasonable values of the parameters from very low $Q^2$ onward ( $Q^2 > 0.1$ GeV$^2$). This result supports the perturbative approach used up to now in model calculations, since it shows, that despite the vicinity of the Landau pole to the hadronic scale, the perturbative expansion is quite convergent and agrees with the non-perturbative results for a wide range of parameters.
In Ref.~\cite{Courtoy:2011mf} the perturbative evolution approach is justified by comparing it to the non-perturbative momentum dependence as determined by the phenomenon of the freezing of the coupling constant, and to analyze the consequences of introducing an effective gluon mass.
\section{Non-perturbative QCD and the Hadron Scale}
\label{sect:NP_scale}
\vspace{.5cm}
The perturbative and non-perturbative approaches can be inferred from the point of view of hadronic models.
We use, as an example, the original bag model, in its most naive description, consisting of a cavity of perturbative vacuum surrounded by non-perturbative vacuum.
The bag model is designed to describe fundamentally static properties, but in QCD all matrix elements must have a scale associated to them as a result of the RGE of the theory. A fundamental step in the development of the use of hadron models for the description of properties at high momentum scales was the assertion that all calculations done in a model should have a RGE scale associated to it \cite{Jaffe:1980ti}. The momentum distribution inside the hadron is only related to the hadronic scale and not to the momentum governing the RGE. Thus a model calculation only gives a boundary condition for the RG evolution as can be seen for example in the LO evolution equation for the moments of the valence quark distribution
\begin{equation}
\langle q_v(Q^2)\rangle_n = \langle q_v(\mu_0^2)\rangle_n \left(\frac{\alpha_s{(Q^2)}}{\alpha_s{(\mu_0^2)}}\right)^{d^n_{NS}},
\label{moments}
\end{equation}
where $d^n_{NS}$ are the anomalous dimensions of the Non Singlet distributions. Inside the bag, the dynamics described by the model is unaffected by the evolution procedure, and the model provides only the expectation value, $\langle q_v(\mu_0^2)\rangle_n$, which is associated with the hadronic scale.
The latter is related to the maximum wavelength at which the structure begins to be unveiled.
This explanation goes over to non-perturbative evolution. The non-perturbative solution of the Dyson--Schwinger equations results in the appearence of an infrared cut-off in the form of a gluon mass which determines the finiteness of the coupling constant in the infrared. The crucial statement is that the gluon mass does not affect the dynamics inside the bag, where perturbative physics is operative and therefore our gluons inside will behave as massless. However, this mass will affect the evolution as we have seen in the case of the coupling constant. The generalization of the coupling constant results to the structure function imply that the LO evolution Eq.~(\ref{moments}) simply changes by incorporating the non-perturbative coupling constant evolution Eq.~(\ref{alphalog}).
\begin{figure}
\begin{center}
\includegraphics[scale= .8]{fig/qval_lo.eps}
\caption{ Left: The running of the effective coupling. The dotted and dashed curves represent the non-perturbative evolution with the parameters used above. The solid curve shows the NNLO evolution with $\Lambda = 250$ MeV.
Right: The evolution of the second moment of the valence quark distribution. The solid curve represents the perturbative LO approximation. }
\label{alpha}
\label{qval}
\end{center}
\end{figure}
%
The non-perturbative results, using the same parameters as before, are quite close to those of the perturbative scheme and therefore we are confident that the latter is a very good approximate description. We note however, that the corresponding hadronic scale, for the sets of parameters chosen, turns out to be slightly smaller than in the perturbative case ($\mu_0^2 \sim 0.1$ GeV$^ 2$), even for small gluon mass $m_0 \sim 0.3$ GeV and small $\rho \sim 1$. One could reach a pure valence scenario at higher $Q^2$ by forcing the parameters but at the price of generating a singularity in the coupling constant in the infrared associated with the specific logarithmic form of the parametrization. We feel that this strong parametrization dependence and the singularity are non physical since the fineteness of the coupling constant in the infrared is a wishful outcome of the non-perturbative analysis. In this sense, the non-perturbative approach seems to favor a scenario where, at the hadronic scale, we have not only valence quarks but also gluons and sea quarks \cite{Scopetta:1997wk,Scopetta:1998sg}. We mean by this statement that to get a scenario with only valence quarks we are forced to very low gluon masses and very small values $\rho$, while a non trivial scenario allows more freedom in the choice of parameters.
\section{The Hadronic Scale from Perturbative Approaches}
\label{sect:soft_evo}
\vspace{.5cm}
Although the perturbative stage of a hard collision is distinct from the non-perturbative regime characterizing the hadron structure, early experimental observations suggest that, in specific kinematical regimes, both {\it the perturbative and non-perturbative stages arise almost ubiquitously}, in the sense that the non-perturbative description follows the perturbative one. In this Section, we discuss an example of perturbative approach in which the transition from perturbative to non-perturbative QCD is clearly identified.
There exists, in DIS processes, a dual description between low-energy and high-energy behavior of a same observable, {\it i.e.} the unpolarized structure functions. Bloom and Gilman observed a connection between the structure function $\nu W_2(\nu, Q^2)$ in the nucleon resonance region and that in the deep inelastic continuum~\cite{Bloom:1970xb,Bloom:1971ye}. The resonance structure function was found to be equivalent to the deep inelastic one, when averaged over the same range in the scaling variable. This concept is known as {\it parton-hadron duality}: the resonances are not a separate entity but are an intrinsic part of the scaling behavior of $\nu W_2$.
The meaning of duality is more intriguing when the equality between resonances and scaling happens at a same scale.
It can be understood as {\it a natural continuation of the perturbative to the non-perturbative representation}.
Bloom--Gilman duality implies a one-to-one correspondence between the behavior of the structure function, $F_2$, for unpolarized electron proton scattering in the resonance region, and in the perturbative QCD regulated scaling region.
In DIS, the relevant kinematical variables are the Bjorken scaling variable, $x=Q^2/2M\nu$ with $M$ being the proton mass and $\nu$
the energy transfer in the lab system, the four-momentum transfer, $Q^2$, and the invariant mass for the proton, $P$, for the virtual photon, $q$, and for the system,
$W^2=(P+q)^2= Q^2\left(1-x\right)/x+M^2$.
For large values of Bjorken $x \geq 0.5$, and $Q^2$ in the multi-GeV$^2$ region, the cross section is dominated by resonance formation, {\it i.e.} $W^2 \leq 5$ GeV$^2$.
While it is impossible to reconstruct the detailed structure of the proton resonances, these remarkably follow the pQCD predictions when averaged over the resonance region.
To answer the question of the nature of a dual description, an option is to focus on purely perturbative analysis from perturbative QCD evolution. Although Bloom--Gilman duality has been known for years, quantitative analyses could be attempted only more recently, having at disposal the extensive, high precision data from Jefferson Lab~\cite{Melnitchouk:2005zr,Liang:2004tj}. Perturbative QCD-based studies~\cite{Liuti:2011rw,Bianchi:2003hi,Liuti:2001qk}, have been presented that include higher-twist contributions or, more generally, the evidence for non-perturbative inserts, which are required to achieve a fully quantitative fit of PDFs, especially at large-$x$. In Ref.~\cite{Courtoy:2013qca}, we discuss the Bloom--Gilman duality from a purely pertubative point of view, by analyzing the scaling behavior of the resonances at the same low-$Q^2$, high-$x$ values as the $F_2$ data from JLab. Our study leads to an analysis of the role of the running coupling constant in the infrared region in tuning the experimental data.
A quantitative definition of \emph{global duality} is accomplished by comparing limited intervals defined according to the experimental data.
Hence, we analyze the scaling results as a theoretical counterpart, or an output of perturbative QCD, in the same kinematical intervals and at the same scale $Q^2$ as the data for $F_2$. It is easily realized that the ratio,
\begin{eqnarray}
R^{\mbox{\tiny exp/th}}(Q^2)&\equiv&\frac{
\int_{x_{\mbox {\tiny min}}}^{x_{\mbox {\tiny max}}} dx\,
F_2^{\mbox {\tiny exp}} (x, Q^2)
}
{\int_{x_{\mbox{\tiny min}}}^{x_{\mbox{\tiny max}}} dx\,
F_2^{\mbox {\tiny th}} (x, Q^2)
}=1\quad,
\label{eq:ratio_1}
\end{eqnarray}
if duality is fulfilled.\footnote{In the analysis of Ref.~\cite{Courtoy:2013qca}, we use, for $F_2^{\mbox {\tiny exp}} $, the data from JLab (Hall C, E94110)~\cite{Liang:2004tj} reanalyzed (binning in $Q^2$ and $x$) as explained in~\cite{Monaghan:2012et} as well as the SLAC data~\cite{Whitlow:1991uw}.}
Duality is violated (the ratio~(\ref{eq:ratio_1}) is not $1$) when considering the fully perturbative expression, and is still violated after corrections by the target mass terms.
One possible explanation for the apparent violation of duality is the lack of accuracy in the Parton Distribution Functions (PDF) parametrizations at large-$x$.\footnote{In our analysis, we use the MSTW08 set at NLO as initial parametrization~\cite{Martin:2009iq}. We have checked that there were no significant discrepancies when using other sets. } Therefore, the behavior of the nucleon structure functions in the
resonance region needs to be addressed in detail
in order to be able to discuss
theoretical predictions in the limit $x \rightarrow 1$. In such a limit, terms containing powers of
$\ln (1-z)$, $z$ being the longitudinal
variable in the evolution equations, that are present in
the Wilson coefficient functions $B_{\mbox{\tiny NS}}^q(z)$ become large and have to be resummed, {\it i.e.} Large-$x$ Resummation (LxR).
Resummation was first introduced by
linking this issue to the definition of the correct kinematical variable that determines the
phase space for real gluon emission
at large $x$. This was found to be $\widetilde{W}^2 = Q^2(1-z)/z$,
instead of $Q^2$~\cite{Amati:1980ch}.
As a result, the argument of the strong coupling constant becomes $z$-dependent~\cite{Roberts:1999gb},
\begin{equation}
\alpha_s(Q^2) \rightarrow \alpha_s\left(Q^2 \frac{(1-z)}{z}\right)\quad.
\end{equation}
In this procedure, however, an ambiguity is introduced, related to the need of continuing
the value of $\alpha_s$
for low values of its argument, {\it i.e.} for $z \rightarrow 1$.
In Ref.~\cite{Courtoy:2013qca}, we have reinterpretated $\alpha_s$ for values of the scale in the infrared region.
To do so, we investigated the effect induced by changing the argument of $\alpha_s$ on the behavior of the $\ln(1-z)$-terms in the convolution with the coefficient function $B_{\mbox{\tiny NS}}$: %
\begin{eqnarray}
F_2^{NS} (x, Q^2)
&=&x q(x,Q^2)+ \frac{\alpha_s}{4\pi} \mathlarger{\sum}_q \mathlarger{\int}_x^1 dz \, B_{\mbox{\tiny NS}}^q(z) \, \frac{x}{z}\, q\left(\frac{x}{z},Q^2\right)\quad,
\label{eq:evo}
\end{eqnarray}
We resum those terms as
\begin{eqnarray}
\ln(1-z)&=&\frac{1}{\alpha_{s, \mbox{\tiny LO}}(Q^2) }\int^{Q^2} d\ln Q^2\, \left[\alpha_{s, \mbox{\tiny LO}}(Q^2 (1-z)) -\alpha_{s, \mbox{\tiny LO}}(Q^2) \right
\equiv \ln_{\mbox{\tiny LxR}}\quad,
\end{eqnarray}
including the complete $z$ dependence of $\alpha_{s, \mbox{\tiny LO}}(\tilde W^2)$ to all logarithms.
Using the ``resummed" $F_2^{\mbox{\tiny theo}} $ in Eq.~(\ref{eq:ratio_1}), the ratio $R$ decreases substantially, even reaching values lower than 1. It is a consequence of the change of the argument of the running coupling constant. At fixed $Q^2$, under integration over $x<z<1$, the scale $Q^2\times(1-z)/z$ is shifted and can reach low values, where the running of the coupling constant starts blowing up. At that stage, our analysis requires non-perturbative information.
In the light of quark-hadron duality, it is necessary to prevent the evolution from enhancing the scaling contribution over the resonances. We define the limit from which non-perturbative effects have to be accounted for by setting a maximum value for the longitudinal momentum fraction, $z_{max}$. Two distinct regions can be studied: the ``running" behavior in $x<z<z_{max}$ and the ``steady" behavior $z_{max}<z<1$.
Our definition of the maximum value for the argument of the running coupling follows from the realization of duality in the resonance region. The value $z_{max}$ is reached at
\begin{eqnarray}
R^{\mbox{\tiny exp/th}}(z_{\mbox{\tiny max}}, Q^2)
&=&\frac{
\mathlarger{\int}_{x_{\mbox {\tiny min}}}^{x_{\mbox {\tiny max}}} dx\,
F_2^{\mbox {\tiny exp}} (x, Q^2)
}
{ \mathlarger{\int}_{x_{\mbox{\tiny min}}}^{x_{\mbox{\tiny max}}} dx\,
F_2^{NS, \mbox{\tiny Resum}} (x, z_{\mbox{\tiny max}}, Q^2)
}
= \frac{I^{\mbox{\tiny exp}}}{I^{ \mbox{\tiny Resum}}}= 1\quad.
\label{eq:ratio_max}
\end{eqnarray}
\begin{figure}[h]
\centering
\includegraphics[scale= .8]{fig/extract_pi_flat_mstw.eps}
\caption{Extraction of $\alpha_s$. See text.
\label{extract}}
\end{figure}
The direct consequence of Eq.~(\ref{eq:ratio_max}) is that duality is realized, within our assumptions, by allowing $\alpha_s$ to run from a minimal scale only. From that minimal scale downward, the coupling constant does not run, it is frozen. This feature is illustrated on Fig.~\ref{extract}. We show the behavior of $\alpha_{s, \mbox{\tiny NLO}}$(scale) in the $\overline{\mbox{MS}}$ scheme and for the same value of $\Lambda_{\overline{\mbox{\scriptsize MS}}, \mbox{\scriptsize MSTW}}^{\mbox{\scriptsize NLO}}=0.402$ GeV used throughout our analysis.
The theoretical errorband correspond to the extreme values of
\begin{equation}
\alpha_{s, \mbox{\tiny NLO}}\left(Q_i^2\frac{(1-z_{max, i})}{z_{max, i }}\right) \qquad,
\end{equation}
$i$ corresponds to the data points.
The determination of the transition scale $Q_0^2$ is probably the main result of our analysis.
Of course, we expect the transition from non-perturbative to perturbative to occur at one unique scale. The discrepancy between the 10 values we have obtained has to be understood as the resulting error propagation. The grey area represents the approximate frozen value of the coupling constant,
\begin{equation}
0.13\leq \frac{\alpha_{s, \mbox{\tiny NLO}} (\mbox{scale}\to 0 \mbox{GeV}^2)}{\pi} \leq 0.18 \quad.
\end{equation}
The solid blue curve represents the (mean value of the) coupling constant obtained from our analysis using inclusive electron scattering data at large $x$. The blue dashed curve represents the exact NLO solution for the running coupling constant in $\overline{\mbox{MS}}$ scheme. The grey area represents the region where the freezing occurs for JLab data, while the hatched area corresponds the freezing region determined from SLAC data. This error band represents the theoretical uncertainty in our analysis.
In the figure we also report values from the extraction using polarized $eP$ scattering data in Ref.~\cite{Deur:2005cf,Deur:2008rf,alexandre}. These values represent the first extraction of an effective coupling in the IR region that was obtained by analyzing the data
relevant for the study of the GDH sum rule. To extract the coupling constant, the $\overline{\mbox{MS}}$ expression of the Bjorken sum rule up to the 5th order in alpha (calculated in the $\overline{\mbox{MS}}$ scheme) was used. The red squares correspond to $\alpha_s$ extracted from Hall B CLAS EG1b, with statistical uncertainties; the orange triangles corresponds to Hall A E94010 / CLAS EG1a
data, the uncertainty here contains both statistics and systematics.
The agreement with our analysis, which is totally independent, is impressive.
We notice, and it is probably one of the most important result of our analysis, that the transition from perturbative to non-perturbative QCD seems to occur around $1$ GeV$^2$.
At that stage, a comparison with fully non-perturbative effective charges and ``modified pQCD" is noteworthy. It is shown in Fig.~\ref{comparisons}.
The grey areas are as in Fig.~\ref{extract} with $\Lambda_{\overline{\mbox{\scriptsize MS}}, \mbox{\scriptsize MSTW}}^{\mbox{\scriptsize NLO}}=0.402$ GeV ; the dashed blue curve is the exact NLO solution with the same $\Lambda$.
The dotted-dashed orange curve corresponds to the result of Ref.~\cite{Fischer:2003rp}, using the version $(b)$ of their
fit with $a=b=1$. The latter analysis was performed in the MOM renormalization scheme. Though the $\beta$ function does not depend on the scheme up to 2 loops, the definition of $\Lambda$ varies from scheme to scheme. The comparison of the results is made possible using the relation,~\cite{Boucaud:1998bq}
\begin{equation}
\Lambda_{\overline{\mbox{\scriptsize MS}}}=\frac{\Lambda_{\mbox{\scriptsize MOM}}}{3.334}\quad,
\end{equation}
leading to the value of $\Lambda_{\overline{\mbox{\scriptsize MS}}}^{\mbox{\scriptsize Ref.~\cite{Fischer:2003rp}}}=(0.71/3.334) $ GeV$\sim 0.21$GeV. The value of $\alpha(0)$ is fixed to $8.915/N_c$.
The red curves are variations of the effective charge of Ref.~\cite{Cornwall:1981zr}, in Eq.~(\ref{alphalog}) with a logarithmic running for the gluon mass described by Eq.~(\ref{eq:mass})
where $(m_0^2, \rho, \Lambda)$ are parameters to be fixed. The solid red curve corresponds to the set $(m_0^2=0.3 $GeV$^2, \rho=1.7, \Lambda=0.25$GeV$)$, the dashed red curve to $(m_0^2=0.5 $GeV$^2, \rho=2., \Lambda=0.25$GeV$)$.
This result is also obtained in the MOM scheme, the value of $\Lambda$ turns out to be similar in both Fischer {\it et al.} and Cornwall's approaches. The cyan curves correspond to two scenarios of the effective charges of Ref.~\cite{Aguilar:2009nf}. Their numerical solution is fitted by a functional form similar to Eq.~(\ref{alphalog}). The 2 sets of parameters, corresponding to $m_0=500$MeV (dashed-dotted curve) and $600$ MeV (medium dashed curve), are then driven by the shape of the numerical solution. They are plotted here with the same $\Lambda_{\mbox{\tiny MOM}}^{n_f=0}=300$ MeV as in the publication, but for $n_f=3$ for sake of comparison. Further investigation on comparison of schemes is needed.
The short dashed green curves corresponds to Shirkov's analytic perturbative QCD to LO~\cite{Shirkov:1997wi} with $\Lambda_{\overline{\mbox{\scriptsize MS}}, \mbox{\scriptsize MSTW}}^{\mbox{\scriptsize LO}}$. The value of $\alpha(0)$ is fixed to $4\pi/\beta_0$. Finally, the pink curve is the freezing value of Ref.~\cite{Mattingly:1992ud}.
Notice that the freezing value for $\alpha_s(Q^2<1$GeV$^2)$ is only constrained by the integral in the resummed version of Eq.~(\ref{eq:evo}): no conclusion can be drawn on its value at $Q^2=0$GeV$^2$. While it is not possible to conclude on the value of $\alpha_s(0)$, we notice that it is possible to find sets of parameters for which the transition from perturbative to non-perturbative QCD occurs around $1$ GeV$^2$.
\begin{figure}[tb]
\centering
\includegraphics[scale= 1.]{fig/compar_pi_mstw_log.eps}
\caption{ Comparison of the effective coupling constant from Ref.~\cite{Courtoy:2013wla}. See text. }
\label{comparisons}
\end{figure}
\section{PDFS at Large-$x$}
\label{sect:largex}
\vspace{.5cm}
As discussed in the previous Section, the apparent violation of duality could be explained by the lack of accuracy in the Parton Distribution Functions (PDF) parametrizations at large-$x$. The remedy is obviously a better understanding and description of the large-$x$ PDFs.
The large-$x$ resummation proposed in Ref.~\cite{Courtoy:2013qca} could be implemented at the PDF level, such that, in a global fit procedure, the large-$x$ region would account for non-perturbative effects, leading to a ``cleaner" functional form for $q(x)$.
An improved treatment of nuclear effects, especially in the resonance region, has already been carried out by the CJ (CTEQ-JLab) collaboration~\cite{Owens:2012bv}. Independent and complementary advancements are needed as uncertainties at large-$x$ highly matter for, {\it e.g.}, search of New Physics' particles.
The impact of PDF uncertainties at large-$x$ on heavy boson production has been studied in Ref.~\cite{Brady:2011hb}, using the CJ PDF set\footnote{The new version is cited but the 2011 version was used.}.
Hadron-hadron collisions involve at least two interacting partons, with momentum fractions $x_1$ and $x_2$, respectively. At fixed center of mass energy $\sqrt{s}$ and boson rapidity $2\,y=\ln\left( \frac{E+p_z}{E-p_z}\right)$ where $E$ and $p_z$ are the boson energy and longitudinal momentum in the hadron center of mass frame, the parton momentum fractions are given (at leading order in the strong coupling constant) by
\begin{eqnarray}
x_{1,2} = \frac{M}{\sqrt{s}} \exp{ (\pm y)}\quad,
\end{eqnarray}
where $M$ is the mass of the produced boson.
At low rapidities the cross sections are relatively insensitive to uncertainties in the large-$x$ PDFs. At larger rapidities, however, there is far greater sensitivity to the large-$x$ behavior, leading to $\approx 15 \%$ uncertainty in the differential cross section for $y_Z = 4$ at the LHC, and for $y_Z = 2.8$ at the Tevatron, which correspond to parton fractions of $x \approx 0.7$.
As for the elusive $W', Z'$ bosons, it is clear now that the increasing of their mass $M$ will directly increase the relevant PDF's $x$ values, so that higher mass bosons will more readily sample the high-$x$ region where the nuclear uncertainties are more prominent. The cross sections will also decrease rapidly with increasing boson mass, so that the effects of the large-$x$ PDF uncertainties will become more significant as the mass increases. The effect on the exclusion limits from, {\it e.g.}, ATLAS~\cite{ATLAS-CONF-2012-129} is important.
\section{Tensor and Scalar Charges}
\label{sect:charges}
\vspace{.5cm}
As we have already stated in these proceedings, the hadronic structure carries important information for high energy observables. Hadronic observables are often related to the manifestation of fundamental processes at the quark level. We here discuss the structural scalar and tensor currents.
\subsection{Relation to New Physics}
\vspace{.5cm}
Non-standard electroweak couplings, studied in the decay of ultracold neutrons~\cite{Bhattacharya:2011qm}, are proposed that are related to hadronic matrix elements. Beyond the well-known weak interactions of the Standard Model, new physics'coupling could be probed in neutron $\beta$-decay. The latter are related to the isovector scalar and axial-vector hadronic matrix elements ~\cite{Weinberg:1958ut},
\begin{subequations}
\begin{eqnarray}
\left \langle p(P_p) \right | \bar{u} d \left | n(P_n)\right\rangle &=& g_S(\Delta^2)\, {\overline u}_p(P_p) u_n(P_n)\quad, \\
\left \langle p(P_p) \right | \bar{u} \sigma_{\mu\nu} d \left | n(P_n)\right\rangle &=& g_T(\Delta^2)\, {\overline u}_p(P_p) \sigma_{\mu\nu}u_n(P_n)+ \ldots\quad,
\end{eqnarray}
\end{subequations}
with $\Delta=P_n-P_p$ and the ellipsis refer to higher order terms.
An analysis of the uncertainties in the spin-independent and spin-dependent elastic scattering cross sections of supersymmetric dark matter particles on protons and neutrons~\cite{Ellis:2008hf} concludes that the largest single uncertainty comes from the spin-independent scattering matrix element $\left \langle N \right | \bar{q} q \left | N\right\rangle$ linked to the $\sigma_{\pi N}$ term.
The ---isoscalar and isovector--- scalar charges and the ---isovector--- tensor charge correspond to the form factors for $\Delta^2=0$, {\it i.e.}\footnote{Where we have dropped the dependence on the renormalization point.}
\begin{subequations}
\begin{eqnarray}
\langle 1\rangle_{\sigma_u-\sigma_d}&=&g_{S}(0)\quad,\\
\langle 1\rangle_{\sigma_u+\sigma_d}&=&\sigma(0)=\frac{\sigma_{\pi N}}{(m_u+m_d)/2}\quad,\label{eq:spn}\\
\langle 1\rangle_{\delta_u-\delta_d}&=&g_{T}(0)\quad.
\label{eq:isovectorTS}
\end{eqnarray}
\end{subequations}
Those matrix elements are not directly accessible through experiments, at least for $\Delta^2=0$ ; exept for the $\sigma_{\pi N}$ related to the form factor $\sigma(2 m_{\pi}^2)$ anlyzed in Ref.~\cite{Pavan:2001wz}.
However, those charges are related to bilocal hadronic matrix elements, defining the PDFs~(\ref{eq:pdf}), through ``sum rules"~\cite{Jaffe:1992ra}.
For each quark flavor, the scalar and axial charges are related to the following forward matrix elements, respectively,
\begin{eqnarray}
\langle 1\rangle_{\sigma_q}(Q^2)&\equiv& \int_0^1 dx \left[ e^q(x,Q^2) +e^{\overline q} (x,Q^2)\right] \quad,\nonumber\\
\Rightarrow M \,e^q(x, Q^2)&=& \int\frac{d\xi^-}{4\pi} e^{i x P^+\xi^-}\left\langle PS\right| {\overline q}(0)\, q(\xi)\left |PS\right \rangle\left |_{\xi^+=\vec{\xi}=0}\right. \quad;
\label{eq:scalarSR}
\\
\nonumber\\
\nonumber\\
\langle 1\rangle_{\delta_q}(Q^2)&\equiv& \delta q (Q^2) = \int_0^1 dx \left[ h_1^q(x,Q^2) - h_1^{\overline q} (x,Q^2)\right] \quad,\nonumber\\
\Rightarrow h_1^q(x, Q^2)&=& \int\frac{d\xi^-}{4\pi} e^{i x P^+\xi^-}\left\langle PS_{\perp}\right| {\overline q}(0)\, i\sigma^{\perp +}\gamma_5q(\xi)\left |PS_{\perp}\right \rangle\left |_{\xi^+=\vec{\xi}=0}\right.\quad,
\label{eq:transversitySR}
\end{eqnarray}
where $Q^2$ is the renormalization scale.
\\
Notice that, contrary to the vector $g_V$ or axial charge $g_A$, there is no proper sum rule associated to the tensor ``charge" (the name in itself is inaccurate).
There are non-vanishing anomalous dimension associated to it and the tensor charge therefore evolves with
the hard scale $Q^2$~\cite{Jaffe:1992ra}.
The dependence on the renormalization scale is given by~\cite{Kodaira:1979ib,Artru:1989zv}, to LO,
\begin{eqnarray}
\delta q(Q^2)&=&\left[ \frac{\alpha_s(Q^2)}{\alpha_s(Q_0^2)}\right]^{-4/27} \delta q(Q_0^2)
\label{eq:evoLO}
\end{eqnarray}
where the exponent comes from, for $n_f=3$, $-2\Delta_T \gamma_{qq}^{(0)}(1)/\beta_0$ with the anomalous dimensions $\Delta_T \gamma_{qq}^{(0)}(n)=\frac{4}{3}\left(\frac{3}{2}-2[\psi(n+1)+\gamma_E]\right)$, with $\psi(n)=d \ln \Gamma(n)/dn$.
the tensor charge has been calculated on
the lattice~\cite{Gockeler:2006zu,Green:2012ud,Bhattacharya:2011qm} and in various
models~\cite{Cloet:2007em,Wakamatsu:2007nc,He:1995gz,Pasquini:2006iv,Gamberg:2001qc},
and it turns out not to be small.
On the other hand, the $\sigma_{\pi N}$ is renormalization point invariant. The sum rule is however related to a singularity and is not of practical use.
While the axial charge is a charge-even
operator, from Eq.~(\ref{eq:transversitySR}), it is evident that the tensor
charge is odd under charge conjugation and, therefore, it does not receive
contributions from $q\bar{q}$ pairs in the sea and is
dominated by valence contributions.
\subsection{Determination of Tensor and Scalar Charges through PDFs}
\vspace{.5cm}
The main experimental access to the tensor and scalar charges is provided thanks to the sum rules Eqs.~(\ref{eq:scalarSR}, \ref{eq:transversitySR}). The knowledge on the PDFs $h_1(x)$ and $e(x)$, up to theoretical limitations, will bring some light on the values of $g_{S/T}$.
Let's start with the transversity PDF.
A comprehensive review of the properties of the transversity distribution
function can be found in Ref.~\cite{Barone:2001sp}.
Transversity $h_1$, as leading-twist collinear PDF, enjoys the same status as
$q$ and $g_1$~\cite{Ralston:1979ys,Jaffe:1992ra}.
The distribution of transversely polarized quarks $q^{\uparrow}$
in a transversely
polarized nucleon $p^{\uparrow}$ (integrated over transverse momentum) can be written as
\begin{equation}
f_{q^{\uparrow}/p^{\uparrow}}(x) = q(x) + {\bf S}\cdot{\bf S}_q \, h_1^q(x) \quad ,
\label{eq:radici-h1dens}
\end{equation}
in which $\bf{S}$ is the nucleon spin and $\bf{S}_q$ the quark spin.
Therefore, transversity can be interpreted as the difference between the
probability\footnote{The probabilistic interpretation is valid in the light-cone gauge.} of finding
a parton (with flavor $q$ and momentum fraction $x$) with
transverse spin parallel and anti-parallel to that of the transversely polarized nucleon.
The only first principle based property on the transversity distribution is the Soffer inequality. Because a probability must be positive, we get the
important
Soffer bound~\cite{Soffer:1995ww},
\begin{equation} 2|h_1^q(x,Q^2)| \leq q(x,Q^2) + g_1^q(x,Q^2) \; ,
\label{eq:radici-soffer}
\end{equation}
which is true at all $Q^2$~\cite{Bourrely:1998bx,Vogelsang:1997ak}. An analogous relation holds for antiquark distributions.
In
spin-$\textstyle{\frac{1}{2}}$ hadrons there is no gluonic function analogous
to transversity. The most important
consequence is that $h_1^q$ for a quark with flavor $q$ does not mix with
gluons in its evolution and it behaves as a non-singlet quantity; this has
been verified up to NLO, where chiral-odd evolution kernels have been studied so
far~\cite{Hayashigaki:1997dn,Kumano:1997qp,Vogelsang:1997ak}.
There are two complementary extractions of the transversity distributions from semi-inclusive processes: the TMD parametrization (also known as Torino fit)~\cite{Anselmino:2008jk,Anselmino:2013vqa} and the collinear extraction (also known as Pavia fit)~\cite{Bacchetta:2011ip,Bacchetta:2012}. The former is based on the TMD framework in which the chiral-odd partner of $h_1(x, k_{\perp})$ is the Collins fragmentation function ; the latter is based on a collinear framework, involving the chiral-odd dihadron fragmentation function, {\it i.e.} $H_1^{\sphericalangle}$. Parameterizations for the dihadron FFs have been obtained independently~\cite{Courtoy:2012ry}. One of the main differences lie in that the collinear extraction does not require the use of a fitting functional form: it is a point-by-point extraction. However, for practical reasons, a statistical study of the transversity PDF has been performed as well. So far, both approaches has found compatible results in the range in $x$ where data exist. However, recent progress on TMD evolution are expected to affect the Torino fit. It is important to notice that the parameterizations are biased by the choice of the fitting functional form. The behavior of the best-fit parametrization is largely unconstrained outside the range of data, leading to confusing results at low and large-$x$ values. This is nicely illustrated by the collinear ---Pavia--- transversity collaboration, on Fig.~\ref{fig:fit_pavia}, where 2 different functional forms, with an equally good $\chi^2/d.o.f.$, have been used.
\begin{figure}[h]
\begin{center}
\includegraphics[width=7cm] {fig/trans_u_full_24_newtorino.eps}
\includegraphics[width=7cm] {fig/trans_d_full_24_newtorino.eps}
\includegraphics[width=7cm] {fig/trans_u_ultra_24_newtorino.eps}
\includegraphics[width=7cm] {fig/trans_d_ultra_24_newtorino.eps}
\end{center}
\caption{Fit of valence collinear transversities at $Q^2=2.4$GeV$^2$. For $u$ and $d$ valence distributions, respectively, left and right columns. The two upper plots correspond to the fits using a flexible parametrization, while the two lower plots correspond to a fit using an extra-flexible parameterization, see Ref.~\cite{Bacchetta:2012}. The red bands represent the standard fit within $1\sigma$ errors, while the green bands stand for a $1\sigma$ Monte Carlo based analysis ($n\times 64\%$ fit of $n$ data replica at $1\sigma$). The light blue bands correspond to the Torino13~\cite{Anselmino:2013vqa} and the full blue curves to the Soffer bound. }
\label{fig:fit_pavia}
\end{figure}
Another independent access to the tensor charge has recently been published using exclusive processes~\cite{Goldstein:2014aja}. Through yet another ``sum rule", the tensor charge corresponds to the first Mellin moment of the chiral-odd $H_T$ generalized parton distribution.
In Fig.~\ref{fig:tenscharge}, we summarize the current status on the tensor charge for up and down quarks. The three extractions are shown at the scale used by each group ; the lattice results are shown for $Q^2=2$ GeV$^2$. In Fig.~\ref{fig:tenscharge_new}, we illustrate the strong dependence on the functional form. More data, especially in the low and large $x$ regions, are needed to constrain the fits. Hopefully, more data from PHENIX are soon to be released. Worth to mention too are the proposals at JLab ---both for CLAS12 and SoLID~\cite{ourprop,ourloi}.
\begin{figure}[htb]
\begin{center}
\includegraphics[height=6.5cm]{fig/tensorcharge_extr.eps}
\end{center}
\caption{
The tensor charge $\delta u$ vs. $\delta d$, computed using the transversity
distributions from TMD collaboration~\cite{Anselmino:2013vqa} (standard) (yellow diamond). The blue circle comes from the chiral-odd GPD $H_T$ sum rule, with the GPD fits of Ref.~\cite{Goldstein:2014aja}.
The purple square corresponds to the standard flexible version of the fit via DiFF~\cite{Bacchetta:2012}, see Fig.~\ref{fig:tenscharge_new} for comparison of the 2 fits' results. The cyan curve corresponds to the lattice result from Ref.~\cite{Green:2012ud} ; the brown curve to Ref.~\cite{Bhattacharya:2013ehc}. }
\label{fig:tenscharge}
\end{figure}
\begin{figure}[htb]
\begin{center}
\includegraphics[width=7.cm]{fig/tensor_ch.eps}
\includegraphics[width=7.cm]{fig/tensor_ch_full.eps}
\end{center}
\caption{{\it Left panel}: tensor charge integrated over data range for the u-quark. {\it Right panel}: tensor charge integrated on the theoretical support $x\in [0,1]$. Respectively for, from 1 to 8, standard rigid, Monte Carlo rigid, standard flexible, Monte Carlo flexible, standard extra-flexible, Monte Carlo extra-flexible of the collinear fit~\cite{Bacchetta:2012} and the fit for $A_0$ and $A_{12}$ asymmetries at Belle combined with single-hadron SIDIS of the TMD fit collaboration~\cite{Anselmino:2013vqa}.}
\label{fig:tenscharge_new}
\end{figure}
As for the scalar charges, they are, though in a more elusive way, related to the twist-3 PDF $e(x)$. See Ref.~\cite{Efremov:2002qh} for a review of the chiral-odd twist-3 PDF. QCD equations of motion allow to decompose the chiral-odd twist-3 distributions into 3 term
\begin{eqnarray}
e^q(x)&=&e_{\mbox{\tiny{loc}}}^q(x)+e_{\mbox{\tiny{tw-3}}}^q(x)+e_{\mbox{\tiny{mass}}}^q(x)\quad.
\end{eqnarray}
The first term comes from the local operator and is related to the pion-nucleon sigma term,
\begin{eqnarray}
e_{\mbox{\tiny{loc}}}^q(x)&=&\frac{1}{2M}\, \int \frac{d\lambda}{2\pi}\, e^{i\lambda x}\langle P| \bar{\psi}_q(0)\psi_q(0)|P\rangle\quad,\nonumber\\
&=&\frac{\delta(x)}{2M}\langle P| \bar{\psi}_q(0)\psi_q(0)|P\rangle\quad;
\end{eqnarray}
the second term is a genuine twist-3 contribution, {\it i.e.} pure quark-gluon interaction term ; while the last term is related to the current quark mass.
Owing to Eq.~(\ref{eq:spn}), $\sigma_{\pi N}$ is related to $e(x=0)$, due to the delta-function singularity.
Furthermore, being a subleading contribution, this twist-3 PDF is hardly known. It is however accessible in single-~\cite{Efremov:2002ut,Gohn:2014zbz} and two-hadron~\cite{silvia_analisi} semi-inclusive DIS off unpolarized targets, through the Beam Spin Asymmetry $A_{LU}$ at CLAS. Just like in the case of transversity extraction, the chiral-odd partner of $e(x)$ is the chiral-odd dihadron fragmentation function, {\it i.e.} $H_1^{\sphericalangle}$. The analysis of the soon-to-be-released data is ongoing~\cite{CourMiraPisa}.
\\
The bounds on $g_S, g_T$, together with the bounds on beta decay couplings, $\tilde{C}_S, \tilde{C}_T$, constrain the new effective couplings, $\epsilon_S, \epsilon_T$, through the matching conditions from a quark-level effective theory to a nucleon-level effective theory~\cite{Bhattacharya:2011qm}
\begin{eqnarray}
\tilde{C}_S&=&g_S \epsilon_S\quad,\nonumber\\
\tilde{C}_T&=&4 g_T \epsilon_T\quad,\nonumber
\end{eqnarray}
The bounds resulting from the phenomenological extractions of the tensor charge~\cite{Bacchetta:2012,Anselmino:2013vqa,Goldstein:2014aja} on the observability of new physics are still huge compared to that the lattice calculations can nowadays achieve. An analysis of the precise projection of these bounds, together with the expected precision of future hadronic structure dedicated experiments is ongoing~\cite{CourGonzLiuti}.
\section{Conclusions}
\vspace{.5cm}
Hadronic physics is the perfect framework to study the intersection between perturbative and non-perturbative QCD. The properties of hadrons reflect the features of the strong interactions and transition from chiral symmetry to confinement.
Non-perturbative QCD provides inputs for perturbative calculation, mainly through {\it initial conditions} for the QCD evolution equations. As described in these proceedings, Parton Distribution Functions in QCD have a scale dependence, through the RGE. Non-perturbative predictions often come from evaluation in models for the proton structure.
This will, in turn, strongly affect the PDF fits: it is called {\it procedural bias} in Ref.~\cite{JimenezDelgado:2012zx} and is a problem known in the hadronic community for years, {\it e.g.}~\cite{Traini:1997jz}.
Besides the uncertainty related to the {\it hadronic scale} and the usual statistical errors, we have mentionned the poor knowledge on PDFs at large values of Bjorken-$x$. The uncertainty on the region $x\to 1$ \cite{Brady:2011hb} will affect the exclusion limit for heavy boson like $Z', W'$.
Finally we have mentionned hadronic matrix elements related to New Physics observables.
The understanding of parton distributions at low energy have important repercussions for the high energy physics.
This is why this plethora of distributions and new information about the hadron structure require to be handled carefully via complementary theoretical approaches.
Therefore, in that sense, QCD must be considered as a whole, from the infrared to the ultraviolet region.
\ack
I am grateful to Ted C. Rogers for priceless discussions about the impact of non-perturbative physics on perturbative QCD. I also thank Simonetta Liuti and Vicente Vento, for their advices and support ; Mart\'in Gonz\'alez Alonso for the explanations about his recent works ; Alexei Prokudin and Stefano Melis for sharing the Torino13 transversity.
This work was funded by the Belgian Fund F.R.S.-FNRS via the contract of Charg\'ee de recherches.
\section*{References}
|
\section{Introduction}
\IEEEPARstart{T}{}HE traditional approach of transmitting an image via a communication channel is to perform compression preceding encryption at the sender side; and to decrypt the cipher-image followed by decompression at the receiving side. However, consider a particular scenario in which Alice needs to transmit an image to Bob but wants to keep the image confidential to an untrusted channel provider Charlie. This implies that Alice should encrypt the image moderately and Charlie has to compress the encrypted image without any knowledge of the cryptographic key. At the receiving side, Bob performs both decompression and decryption to reconstruct the original image.
Some works for compressing encrypted images have been reported in recent years. A scheme for compressing encrypted images using a 2-D source model and LDPC codes was developed in \cite{schonberg2006compression}. It is based on the finding that encrypted data are as compressible as unencrypted ones by considering the problem as distributed source coding. The lossless compression of encrypted grayscale and color images has been presented in \cite{lazzeretti2008lossless}, by decomposing the image pixels into bit-planes. By applying the approach of \cite{johnson2004compressing} to the prediction error domain, a better lossless compression performance on the encrypted grayscale and color images is achieved \cite{anil2008distributed}. A progressive compression approach for processing an encrypted image has been suggested, in which the decoder needs to study the local statistics of a low-resolution image and then decodes the next resolution level \cite{liu2010efficient}. Meanwhile, the lossy compression of encrypted images was also studied to achieve higher compression ratios \cite{schonberg2008toward, johnson2004compressing, zhang2011lossy, zhang2012scalable, zhang2013compression, zhou2013designing}. For example, based on the results of \cite{johnson2004compressing}, a practical model for compressing encrypted binary image has been developed in \cite{schonberg2008toward}. Zhang proposed a novel scheme for the lossy compression of an encrypted image at a flexible compression ratio \cite{zhang2011lossy}, in which a pseudorandom permutation is used to encrypt the plain-image. Making use of the process of masking the original pixel values by a modulo-256 addition with pseudorandom numbers, Zhang \emph{et al.} further proposed a scheme for the scalable coding of encrypted images \cite{zhang2012scalable}. In \cite{zhang2013compression}, the compression is performed on an encrypted image with multi-layer decomposition. Zhou \emph{et al.} designed an efficient encryption-then-compression scheme for images via error clustering, in which both lossless and lossy compressions were considered \cite{zhou2013designing}. The above-mentioned approaches of compressing encrypted images are not suitable for high packet loss transmission in non-feedback systems, since the resultant coded streams have substantially unequal importance such that the loss of some codewords may cause severe error propagation and results in unsatisfactory decoded result.
Multiple description coding is a common approach to deal with packet loss during transmission. In general, a multiple description coder generates two or more sub-streams referred to as descriptions. The packets of each description are transmitted over multiple disjoint paths. After receiving each description, the decoder is able to perform a low-quality reconstruction. If all the descriptions have been received, the reconstruction quality is the best. Such a protocol allows a channel with network congestion or packet loss to perform the decoding at the expense of reconstruction quality. Multiple description coding of natural images has been extensively studied in \cite{servetto2000multiple, yang2007robust, li2011balanced, deng2012robust}, where spatial correlations are often eliminated by using sparse transforms like DWT. However, they are not suitable for encrypted images since sparse transforms are nearly ineffective on encrypted images due to the low correlation between the pixels. A multiple description coder especially designed for encrypted images is rarely reported so far.
Consider the scenario that Alice needs the semi-trusted channel coder Charlie to transmit an encrypted image to Bob. When a high packet loss is encountered in the channel between Charlie and Bob, Charlie should first encode the encrypted image for error control. This motivates us to explore a multiple description coder aiming at the robust coding of encrypted images. In this work, we design such a coder based on compressive sensing (CS) with a structurally random matrix (SRM). The proposed coder is comprised of three parts: permutation-based encryption by Alice, encoding using structural matrix (SM) by Charlie, and joint decryption and decoding by Bob. In particular, Alice first encrypts an image using globally random permutation and then sends the encrypted image to the semi-trusted channel encoder Charlie who samples the encrypted image using a structural matrix. Through a channel with high packet loss, Bob receives the compressive measurements and reconstructs the original image by joint decryption and decoding. Moreover, we discuss the relationship between our approach and existing algorithms and describe two other cryptographic applications of SRM. In the performance evaluation, we explore the relationship between packet loss rate and sampling rate and then introduce a feasible quantization approach to the compressive measurements of encrypted images. Finally, we investigate the robustness of the proposed coder at different parameter settings. It is verified that the proposed coder can be regarded as an efficient multiple description coder with a number of descriptions against packet loss.
The rest of this paper is organized as follows. Section II is a brief review of the theory of CS using SRM. In Section III, the robust coding of encrypted images based on CS with SM is proposed. Further discussions can be found in Section IV while the performance evaluation is reported in Section V. Finally, we conclude the paper with some remarks in Section VI.
\vspace{-0.05in}
\section{Compressive Sensing by Structurally Random Matrix}
The fundamental Shannon/Nyquist sampling theory is widely-accepted as the keystone in signal acquisition and reconstruction. It governs the sampling process from the perspective of signal bandwidth. Nevertheless, the number of required measurements can be so large that the storage becomes unbearable and the acquisition time can be very long. Compressive sensing \cite{candes2006robust, donoho2006compressed} is a new sampling theory which allows the exact recovery of a sparse signal from a few linear projections lower than the Nyquist rate. The underlying property of CS is the sparsity of interest. A signal ${\bf{x}}$ of length $N$ is said to be $K$-sparse or compressible if it can be well approximated using only $K \ll N$ coefficients over some sparsifying basis ${\bf{\Psi }}$ as follows
\begin{equation}
{\bf{x = \Psi s}},
\end{equation}
where ${\bf{s}}$ is the transform coefficient vector that contains at most $K$ significant nonzero entries. Compressive sensing theory indicates that ${\bf{x}}$ can be acquired by the following random measurement
\begin{equation}
{\bf{y = \Phi x}},
\end{equation}
where ${\bf{\Phi }}$ is a $M \times N$ ($M < N$) random measurement matrix and ${\bf{y}}$ represents the measurement coefficient vector. ${\bf{x}}$ can be faithfully recovered from only $M = {\rm \mathcal{O}}\left( {K\log N} \right)$ measurements through ${l_1}$-minimization
\begin{equation}
\min {\kern 1pt} {\kern 1pt} {\left\| {\bf{s}} \right\|_1}\; \textrm{s.t.} \;{\bf{y = \Phi \Psi s}},\;\\
\end{equation}
where the measurement matrix ${\bf{\Phi }}$ should be highly incoherent with the sparsifying basis ${\bf{\Psi }}$.
The design of an efficient measurement matrix is still a big challenge in CS. Do \emph{et al.} \cite{do2012fast} introduced a fast and efficient measurement matrix for practical CS. The matrix is called a structurally random matrix (SRM), which, in many aspects, outperforms the existing popular sensing matrices such as Gaussian, Bernoulli and Fourier matrices \cite{candes2006near, mendelson2008uniform, candes2007sparsity}. Gaussian and Bernoulli matrices require high computation complexity and huge memory buffering due to their completely unstructured nature while Fourier matrix works well only if the sparsifying basis is an identity matrix. Do \emph{et al.} also pointed out that SRM possesses the following features: optimal or near-optimal sensing performance; universality; low complexity; hardware/optical implementation friendless. In particular, it is defined as a product of three matrices
\begin{equation}
{\bf{\Phi }} = \sqrt {\frac{N}{M}} {\bf{DFR}}
\end{equation}
where ${\bf{R}} \in {\mathbb{R}^{N \times N}}$ is either a uniform random permutation matrix or a diagonal random matrix whose diagonal entries are Bernoulli random variables. ${\bf{F}} \in {\mathbb{R}^{N \times N}}$ represents an orthonormal matrix that is selected among popular fast computable transforms such as Fast Fourier Transform (FFT), Discrete Cosine Transform (DCT) and Walsh-Hadamard Transform (WHT). ${\bf{D}} \in {\mathbb{R}^{N \times N}}$ is a subsampling operator selecting a random subset of rows of the matrix ${\bf{FR}}$. Interested readers can refer to \cite{do2012fast} for more details on SRM.
\vspace{-0.05in}
\section{Robust Coding of Encrypted Image via Structural Matrix}
Compressing encrypted images is a big challenge due to the fact that an effective encryption algorithm must have already removed or lowered the correlation among neighbouring image pixels to increase the entropy. However, classical image compression schemes like JPEG 2000 always make use of the high correlation and non-uniformity of image pixels. Some lightweight encryption techniques only permute the pixels or mask the pixel values by a keystream. As a result, the encrypted image may still be compressed to certain extent by leveraging some particular coding techniques \cite{johnson2004compressing, schonberg2006compression, lazzeretti2008lossless, anil2008distributed, liu2010efficient, schonberg2008toward, zhang2011lossy, zhang2012scalable, zhang2013compression, zhou2013designing}. The lightweight encryption schemes are usually not secure enough, but they are employed in some specific application scenarios. The proposed scheme does not aim at improving the compression performance on encrypted images but focuses on designing a robust coder for the transmission of encrypted images over a channel with high packet loss rate.
The proposed coder is based on SRM. The basic idea is to split the measurement matrix ${\bf{\Phi }} = \sqrt {{N \mathord{\left/
{\vphantom {N M}} \right. \kern-\nulldelimiterspace} M}} {\bf{DFR}}$ in (4) into two matrices: the matrix ${\bf{R}}$ and the matrix $\sqrt {{N \mathord{\left/
{\vphantom {N M}} \right. \kern-\nulldelimiterspace} M}} {\bf{DF}}$. ${\bf{R}}$ is a random permutation matrix which can serve as a lightweight encryption tool while $\sqrt {{N \mathord{\left/
{\vphantom {N M}} \right. \kern-\nulldelimiterspace} M}} {\bf{DF}}$ can be considered as a new measurement matrix in the proposed coder. First, Alice encrypts an image using ${\bf{R}}$ and then sends the encrypted image to the channel coder Charlie who samples the encrypted image using $\sqrt {{N \mathord{\left/ {\vphantom {N M}} \right. \kern-\nulldelimiterspace} M}} {\bf{DF}}$. Through a high packet loss channel, Bob receives the compressive measurements and reconstructs the original image by joint decryption and decoding using $\sqrt {{N \mathord{\left/ {\vphantom {N M}} \right. \kern-\nulldelimiterspace} M}} {\bf{DFR}}$, as illustrated in Fig. 1. The random permutation ${\bf{R}}$ is constructed from a secret seed known to both Alice and Bob.
The robust coding of encrypted images by structural matrices is composed of three steps: permutation-based encryption by Alice, encoding using structural matrix by Charlie, and joint decryption and decoding by Bob.
\begin{figure}[th]\centering
\includegraphics[width=8 cm]{fig1.eps}\\
\caption{A block diagram of the proposed coder.}
\label{fig1}
\end{figure}
\subsection{Permutation-based Encryption by Alice}
The encrypted image is obtained by applying random spatial permutation on the image. Alice converts the original image ${\bf{X}}$ of size ${N_1} \times {N_2}$ into a vector ${\bf{x}}$ with length $N = {N_1} \times {N_2}$. Then she encrypts ${\bf{x}}$ to the cipher sequence ${{\bf{x}}_{en}}$ by applying a random permutation matrix ${\bf{R}} \in {\mathbb{R}^{N \times N}}$, governed by
\begin{equation}
{{\bf{x}}_{en}} = {\bf{Rx}}.
\end{equation}
${{\bf{x}}_{en}}$ is rearranged into a 2-D cipher image ${{\bf{X}}_{en}}$, which is then sent to Charlie who obtains the encrypted sequence ${{\bf{x}}_{en}}$ by arranging ${{\bf{X}}_{en}}$. The conversion between vector and matrix is known to both Alice and Charlie. The random permutation matrix ${\bf{R}}$ is a binary matrix in which each row or column has exactly one 1 and the rest are all zero. It is generated by a pseudo-random generator with initial random seed shared between Alice and Bob. The reader may refer to \cite{mitra2006new, zhang2014image} for more illustrations on the encryption methods based on permutation matrix. It should be noticed that permutation-based decryption is performed by multiplying the cipher image with the inverse permutation matrix. Interestingly, it is not necessary to invert the matrix since the inverse matrix is obtained by transposing the permutation matrix itself, i.e., ${{\bf{R}}^{ - 1}} = {{\bf{R}}^T}$. The key space is $N!$ so that it is not likely for Charlie to launch a brute force search when $N$ is sufficiently large. Permutation-based encryption cannot hide the statistical information of the original image due to its unaltered histogram. In spite of this, it can still be employed in applications where high secrecy is not a must.
\subsection{Encoding using Structural Matrix by Charlie}
After the encrypted image has been received, Charlie constructs a special measurement matrix to sample it. This matrix is tailored to the encrypted image and is called structural matrix (SM). It is governed by
\begin{equation}
{\bf{A}} = \sqrt {\frac{N}{M}} {\bf{DF}},
\end{equation}
where ${\bf{D}}$ and ${\bf{F}}$ are as described in (4). Encoding using SM is expressed as
\begin{equation}
{\bf{y = A}}{{\bf{x}}_{en}}.
\end{equation}
Obviously, SM is derived from SRM due to the fact that ${\bf{y = A}}{{\bf{x}}_{en}} = \sqrt {{N \mathord{\left/ {\vphantom {N M}} \right. \kern-\nulldelimiterspace} M}} {\bf{DF}}{{\bf{x}}_{en}} = \sqrt {{N \mathord{\left/ {\vphantom {N M}} \right. \kern-\nulldelimiterspace} M}} {\bf{DFRx}} = {\bf{\Phi x}}$. The scenario that SM is applied for permuted or encrypted images is the same as that SRM is employed for spatial images. Structural matrix is expediently selected among some popular computable matrices such as FFT, SCT, WHT or their block diagonal versions. The $M$ rows are extracted at random from SM. These matrices have stable structures like SRM and they outperform Gaussian and Bernoulli matrices in terms of computational complexity and memory requirement. It can be easily inferred that the performance of SM measuring the encrypted image is the same as that of SRM sampling the original image. It has been mathematically proved in \cite{do2012fast} that entries of ${\bf{AR\Psi }}$ asymptotically form a normal distribution $\mathcal{N}\left( {0,{\sigma ^2}} \right)$, where ${\bf{\Psi }}$ is an arbitrary orthonormal matrix and ${\sigma ^2} \le {\rm \mathcal{O}}\left( {\frac{1}{\mathcal{N}}} \right)$, under some mild assumptions: ${\bf{F}}$ is an unit-row matrix whose entries have absolute magnitude in the order of ${\sigma ^2} \le {\rm \mathcal{O}}\left( {\frac{1}{\mathcal{N}}} \right)$ and the sum of entries in each row is equal to zero; ${\bf{\Psi }}$ is an unit-norm column matrix with entries having maximal absolute magnitude in the order of $\mathcal{O}\left( 1 \right)$ and the average sum of entries in each column in the order of ${\sigma ^2} \le {\rm \mathcal{O}}\left( {\frac{1}{\mathcal{N}}} \right)$. The entries in each row of ${\bf{F}}$ and each column of ${\bf{\Psi }}$ are not all equal. Do \emph{et al.} also found that SRM supports block-based models with high incoherence between ${\bf{FR}}$ and ${\bf{\Psi }}$. It should be noticed that the randomization ${\bf{D}}$ can induce a new application scenario, which will be described later.
\subsection{Joint decryption and decoding by Bob}
At the receiving side, Bob obtains the compressive measurements ${\bf{y}}$ and applies joint decryption and decoding to recover the original image using the following algorithm:
\begin{equation}
\min {\kern 1pt} {\kern 1pt} {\left\| {\bf{s}} \right\|_1}\; \textrm{s.t.} \;{\bf{y = AR\Psi s}} = \sqrt {\frac{N}{M}} {\bf{DFR\Psi s}}\;\\
\end{equation}
As a result, ${\bf{x = \Psi s}}$. The recovery criterion has been stated in \cite{do2012fast}: with a probability of at least $1 - \delta $, the sensing framework using SRM can exactly recover $K$-sparse signals if $M \ge {\rm \mathcal{O}}\left( {\frac{N}{B}K{{\log }^2}\frac{N}{\delta }} \right)$, where $B$ is the block size. Theoretically, this guarantees the capability of SM in encoding the encrypted image.
\vspace{-0.05in}
\section{Further Discussions}
In some references \cite{han2010image, wu2011multivariate, zhang2012image, gao2013image}, CS was applied for natural image coding but this is not an appropriate approach in terms of compression efficiency \cite{ goyal2008compressive}. Nevertheless, in view of the robustness property of multiple description coder, CS can be a good candidate \cite{gao2010robust, liu2011new, deng2012robust}. A representative work was presented by Deng \emph{et al.} in \cite{deng2012robust}, in which the compressive measurements can be viewed as a number of descriptions mainly because of their \emph{democracy} properties. If the measurement matrix follows the Gaussian distribution, each CS measurement possesses a similar amount of information of the original signal \cite{davenport2009simple}. Specifically, the sampling is performed on the frequency coefficients generated by two-dimensional DWT and at the decoding side, two different recovery algorithms are developed for the low-frequency and high-frequency subbands, respectively, by fully exploiting the intra-scale and inter-scale correlation of multiscale DWT. Although experimental results showed that this CS-based codec is much more robust for lossy channels in comparison with existing CS-based coding schemes \cite{deng2012robust}, it is not suitable for processing encrypted images. This is because the efficiency of sparse transforms like DWT mainly depends on strong correlation between pixels, which must be weakened by the encryption process, even if a lightweight one is employed.
CS-based compression of encrypted image has been explored in only two references \cite{kumar2009lossy, zhang2011compressing}, both of which aimed at the linear transformation encryption operations. Both coders adopt the block-to-block structure which possesses a straightforward advantage, i.e., parallel CS encoding and decoding. Unfortunately, such a block encryption manner suffers from three drawbacks. Firstly, individual block operation makes the cipher more insecure than global image transform. In order to enhance the security, different blocks may be endowed with different keys and more keys need to be transmitted. Secondly, a plain image is divided into a number of non-overlapping blocks having different statistical features and unequal significance. When these blocks are individually sampled, the measurements have unequal significance. As a result, both coders cannot be considered as efficient multiple description coders. Thirdly, blocking artifact cannot be avoided. In addition, a random matrix is chosen as the measurement matrix. In practical sensing applications, this is costly as very high computational complexity and huge memory buffering are required due to the completely unstructured nature of the matrix \cite{candes2007sparsity}. The proposed coder does not suffer from the above drawbacks. Global permutation is a common lightweight image encryption technique which is more secure than individual block permutation. The random permutation ${\bf{R}}$ relocates all the pixels globally. It destroys the image structure and converts a meaningful image into one look like white noise \cite{do2012fast}. The structural matrix ${\bf{A}}$ in sampling the permuted image supports block processing, meaning that parallel CS encoding can be applied. ${\bf{R}}$ disperses the energy of the whole image and ${\bf{F}}$ further spreads the energy over all the measurements. Consequently, the sampled measurements obtained by SM roughly have the same significance. The proposed coder is a multiple description coder with a number of descriptions whose capability in resisting against packet loss is verified in the next section. There is no blocking artifact as a unified decoder is used to reconstruct the whole image. Compared with random matrix, SM facilitates fast computation and low-complexity electronic or optical implementation.
It is worth mentioning that SRM also induces two other applications related to coding and encryption due to the randomness of ${\bf{R}}$ and ${\bf{D}}$. The first application is illustrated in Fig. 2(a). Alice still permutes the image with ${\bf{R}}$ while Charlie can further encrypt the permuted image with ${\bf{DF}}$. This is because the matrix ${\bf{D}}$ is a random selection operation which can serve as a secret key shared between Charlie and Bob. Another application is the direct encryption by Alice using ${\bf{DFR}}$, as shown in Fig. 2(b). Both applications can be considered as joint coding and encryption schemes. The size of the key space due to ${\bf{D}}$ is given by the combinatorial number $\left( \begin{array}{l}M\\N\end{array} \right)$. It seems that the current size of key space upgraded as $N! + \left( \begin{array}{l}M\\N\end{array} \right)$ is sufficiently large to resist brute-force attack. Unfortunately, the encryption schemes based on CS with SRM is probably insecure against some potential attacks such as known-plaintext attack and chosen-plaintext attack due to its linearity \cite{rachlin2008secrecy}. As a consequence, the security level of CS needs to be analyzed. For example, a low-complexity multiclass encryption scheme has been designed in \cite{cambareri2013low1, cambareri2013low2}, which possesses strong resistance against known-plaintext attacks.
\begin{figure}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.85\linewidth}
\includegraphics[width=\textwidth]{fig2a.eps}
\end{minipage}}\\
\subfigure[]{
\begin{minipage}[t]{0.85\linewidth}
\includegraphics[width=\textwidth]{fig2b.eps}
\end{minipage}}
\caption{\small Two other applications of SRM.}
\label{fig2}
\end{figure}
\begin{figure}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.9\linewidth}
\includegraphics[width=\textwidth]{fig3a.eps}
\end{minipage}}\\
\subfigure[]{
\begin{minipage}[t]{0.9\linewidth}
\includegraphics[width=\textwidth]{fig3b.eps}
\end{minipage}}
\caption{\small PSNRs of the reconstructed images with respect to (a) PLR; (b) SR.}
\label{fig3}
\end{figure}
\begin{figure*}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig4a.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig4b.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig4c.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig4d.eps}
\end{minipage}}
\caption{\small Histograms of the encoded images for the cases: (a) Lena, SR=0.8, BDCT32, PLR=0.05; (b) Peppers, SR=0.6, BWHT32, PLR=0.10; (c) Boat, SR=0.8, BDCT32, PLR=0.15; (d) Goldhill, SR=0.6, BWHT32, PLR=0.20.}
\label{fig4}
\end{figure*}
\begin{figure}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.9\linewidth}
\includegraphics[width=\textwidth]{fig5a.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.9\linewidth}
\includegraphics[width=\textwidth]{fig5b.eps}
\end{minipage}}
\caption{\small The values of $\gamma $ versus PLR for (a) SR=0.6; (b) SR=0.8.}
\label{fig5}
\end{figure}
\begin{figure}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.9\linewidth}
\includegraphics[width=\textwidth]{fig6a.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.9\linewidth}
\includegraphics[width=\textwidth]{fig6b.eps}
\end{minipage}}
\caption{\small Rate-Distortion performance of the quantization for (a) SR=0.6, BDCT32; (b) SR=0.6, BWHT32.}
\label{fig6}
\end{figure}
\vspace{-0.05in}
\section{Performance Evaluation}
Our simulation settings are similar to those using SRM \cite{do2012fast}. Four natural images of size 512$\times$512 including \emph{Lena}, \emph{Peppers}, \emph{Boat} and \emph{Goldhill} are used for testing. The sparsifying basis ${\bf{\Psi }}$ is Daubechies 9/7 wavelet transform. The reconstruction algorithm is GRSR in \cite{figueiredo2007gradient}. ${\bf{R}}$ and ${\bf{D}}$ are generated using MATLAB commands and ${\bf{F}}$ is chosen as block diagonal DCT (BDCT) and block diagonal WHT (BWHT). The packet size is set to 100 unless specified. We first explore the relationship between packet loss rate and sampling rate and then describe a feasible quantization approach for the compressive measurements of encrypted images. Finally, the robustness of the proposed coder at different parameter settings is investigated.
\subsection{Relationship between packet loss rate and sampling rate}
The compressive measurements ${\bf{y}}$ of length $M$ can be partitioned, at equal intervals, into a number of packets. Each packet carries a similar amount of information of the original image since all the measurements have roughly equal importance. If a packet contains $m$ measurements, there are $\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil $ packets in total. Lost packets always occur randomly and Bob will update ${\bf{D}}$ according to the received packets. We denote packet loss rate as PLR which can be up to 30\% in real cases \cite{zhao2003understanding}. The sampling rate (SR) is defined as $SR = {M \mathord{\left/ {\vphantom {M N}} \right. \kern-\nulldelimiterspace} N}$. For example, if $M = 157290$ and $m = 100$, then $SR = {{157290} \mathord{\left/ {\vphantom {{157290} {{{512}^2}}}} \right. \kern-\nulldelimiterspace} {{{512}^2}}} = 0.60$ and the number of packets is $\left\lceil {{{157290} \mathord{\left/ {\vphantom {{157290} {100}}} \right. \kern-\nulldelimiterspace} {100}}} \right\rceil = 1573$. If $PLR = 0.20$, the number of lost packets is $1573 \times 0.2 \buildrel\textstyle.\over= 315$ and the number of received packets is 1258. In other words, Charlie sends ${512^2}$ measurements and Bob receives about 125800 measurements among them. This is similar to the case that the sampling rate is changed to $SR' = {{125800} \mathord{\left/ {\vphantom {{125800} {{{512}^2}}}} \right. \kern-\nulldelimiterspace} {{{512}^2}}} \buildrel\textstyle.\over= 0.48$. In fact, this equivalence is reasonable due to the roughly equal importance of the measurements. This example inspires us a relationship between SR and PLR.
In general, for a given $SR = \alpha $ $\left( {0 < \alpha < 1} \right)$, $PLR = \beta $ $\left( {0 \le \beta \le 0.3} \right)$ is basically equivalent to $SR = \alpha \left( {1 - \beta } \right)$. This can be verified in Fig. 3, where BDCT32 and BWHT32, corresponding to the solid line and the dashed line, respectively, mean that each sub-matrix in the diagonal of ${\bf{F}}$ has a size of 32$\times$32. It can be observed that the effects of BDCT and BWHT are consistent since each pair of solid and dashed lines coincides with each other while other conditions are identical. The value of SR is set as $SR{\rm{ = }}0.6$ in Fig. 3(a). $PLR = \beta $ in Fig. 3(a) corresponds to $SR = 0.6 \times \left( {1 - \beta } \right)$ in Fig. 3(b). A comparison between Fig. 3(a) and Fig. 3(b) shows that the former PSNR roughly coincides with the latter one. Both starting points have the same PSNR value, i.e., $PLR = 0$ in Fig. 3(a) and $SR{\rm{ = }}0.6$ in Fig. 3(b). However, with the increase of PLR and the reduction of SR, the PSNR value of the former is sightly lower than that of the latter. There are three factors causing this difference: (i) Weak correlations exist between adjacent measurements. The amount of information of the whole packet containing $m$ successive measurements is gracefully greater than that provided by the $m$ randomly-sampled measurements; (ii) After packing the measurements, the number of measurements $m'$ in the last packet is less than $m$ as long as $M$ is not divisible by $m$. The last packet will not be lost with high probability $\left( {1 - \beta } \right)$ such that the actual $SR = \alpha {{\left( {m\left( {\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil \left( {1 - \beta } \right) - 1} \right) + m'} \right)} \mathord{\left/ {\vphantom {{\left( {m\left( {\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil \left( {1 - \beta } \right) - 1} \right) + m'} \right)} M}} \right. \kern-\nulldelimiterspace} M} < \alpha \left( {1 - \beta } \right)$. (iii) The rounding effect of $\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil \beta $ possibly results in the actual $PLR = {{round\left( {\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil \cdot \beta } \right)} \mathord{\left/ {\vphantom {{\textrm{round}\left( {\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil \cdot \beta } \right)} {\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil }}} \right. \kern-\nulldelimiterspace} {\left\lceil {{M \mathord{\left/ {\vphantom {M m}} \right. \kern-\nulldelimiterspace} m}} \right\rceil }} > \beta $. Revealing such a connection of PLR and SR helps to adjust the SR according to the PLR in real-time transmission. Bob distinguishes the PLR according to the received packets and then feeds back to Charlie who adjusts the SR to guarantee a certain PSNR value for the image received by Bob.
\begin{figure*}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig7a.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig7b.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig7c.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig7d.eps}
\end{minipage}}
\caption{\small The reconstructed images and their PSNR values under SR=0.8: (a) PSNR=35.7965, PLR=0.2, BWHT32; (b) PSNR=33.8944, PLR=0.3, BWHT32; (c) PSNR=35.2942, PLR=0.2, BDCT32; (d) PSNR=33.2762, PLR=0.3, BDCT32.}
\label{fig7}
\end{figure*}
\begin{figure*}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig8a.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig8b.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig8c.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig8d.eps}
\end{minipage}}
\caption{\small The reconstructed images and their PSNR values under SR=0.5: (a) PSNR=32.6432, PLR=0.2, BWHT32; (b) PSNR=31.2287, PLR=0.3, BWHT32;
(c) PSNR=32.0629, PLR=0.2, BDCT32; (d) PSNR=30.7208, PLR=0.3, BDCT32.}
\label{fig8}
\end{figure*}
\begin{figure*}[th]
\centering
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig9a.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig9b.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig9c.eps}
\end{minipage}}
\subfigure[]{
\begin{minipage}[t]{0.22\linewidth}
\includegraphics[width=\textwidth]{fig9d.eps}
\end{minipage}}
\caption{\small The reconstructed images and their PSNR values under SR=0.2: (a) PSNR=26.2722, PLR=0.2, BWHT32; (b) PSNR=25.4766, PLR=0.3, BWHT32;
(c) PSNR=26.1835, PLR=0.2, BDCT32; (d) PSNR=24.9015, PLR=0.3, BDCT32.}
\label{fig9}
\end{figure*}
\begin{table*}[tb]
\caption{PSNR versus Round-off and Without Round-off (Lena, SR=0.6, BDCT32).}
\centering
\begin{tabular}{l c c c c c c c}
\hline
$PLR$ & 0 & 0.05 & 0.10 & 0.15 & 0.20 & 0.25 & 0.30 \\
\hline
Round-off & 37.18 & 36.37 & 35.54 & 34.59 & 33.79 & 33.16& 32.15 \\
Without round-off & 37.22& 36.37 &35.66 &34.61 &34.00 &33.25& 32.34 \\
Difference & 0.04 & 0.00 & 0.12 & 0.02 & 0.21 & 0.09 & 0.19 \\
\hline
\end{tabular}
\label{tab1}
\end{table*}
\begin{table*}[tb]
\caption{PSNR versus Round-off and Without Round-off (Lena, SR=0.8, BWHT32).}
\centering
\begin{tabular}{l c c c c c c c}
\hline
$PLR$ & 0 & 0.05 & 0.10 & 0.15 & 0.20 & 0.25 & 0.30 \\
\hline
Round-off & 40.96 &39.27 &38.21& 36.82 &35.70 &34.69 &33.96 \\
Without round-off & 41.01& 39.54& 38.34 &36.82 &35.72& 35.02 &34.00 \\
Difference & 0.05& 0.27 & 0.13 & 0.00 & 0.02 & 0.33 & 0.04 \\
\hline
\end{tabular}
\label{tab2}
\end{table*}
\begin{figure}[th]\centering
\includegraphics[width=8 cm]{fig10.eps}\\
\caption{PSNR versus packet size when SR=0.6 and PLR=0.3.}
\label{fig10}
\end{figure}
\begin{figure}[th]\centering
\includegraphics[width=8 cm]{fig11.eps}\\
\caption{PSNR versus block size of SM when SR=0.6 and PLR=0.05.}
\label{fig11}
\end{figure}
\subsection{Quantization of Compressive Measurements of Cipher Image}
When the compressive measurements are transmitted over a communication channel, they need to be efficiently quantized and encoded. Therefore, the measurements' statistics are required and an optimal quantizer should be tailored to the measurements for minimizing the amount of distortion during reconstruction. The statistical distribution of compressive measurements obtained by SRM has been well studied \cite{haimi2013distribution}. It has been pointed out that the encryption performed by a random permutation on the pixel indices makes the measurements suitable for quantization by causing the measurements' distribution roughly normal. The measurements obtained by applying SM to the encrypted image approximately yield a Gaussian distribution. This is also observed in Fig. 4, which depicts the histograms of various encoded images in different cases.
A uniform scalar quantization is employed to round each entity of ${\bf{y}}$ to the nearest integer. The difference in distortion caused by the round-off is extremely subtle, as shown in Tables I and II. Moreover, we can observe from Fig. 4 that the measurement values roughly lie between -150 and 150. The farther the measurement value deviates from zero, the fewer the number of measurements are required. Our quantization method only reserves and rounds the values located within the interval $[ - 127.5,127.5)$. Others are discarded due to two reasons: (i) The discarded measurements make up only a low proportion, marked as $\gamma $, of the whole measurements. Figure 5 lists the values of $\gamma $ at different parameter settings. $\gamma $ is basically smaller than 0.0055, which implies that either the PLR rises slightly to $PLR = \beta + \gamma $ or the SR drops a small portion $\alpha \gamma $ by the reason of the approximately equal importance among the measurements; (ii) The reserved measurement values can be one-to-one mapped to the interval $[0,255]$ through adding 128 to every value. The integers in $[0,255]$ not only can be fully represented by 8-bit numbers, but also match with the common-adopted 256 grayscales in the images. After the encoding process is completed, an image can still be stored in 8-bit format, which leads to great convenience in practical usage.
The quantization distortion is caused by two factors: the decimal round-off and the proportion of discarded measurements. The first factor is insignificant, as justified by the data listed in Tables I and II while the second one is the same because $\gamma $ is basically smaller than 0.0055. It can also be justified by the rate-distortion curves plotted in Fig. 6, in which the dashed and solid lines correspond to cases with and without quantization, respectively. These two lines are almost identical and they indicate that the proposed quantization method works well.
\subsection{Robustness}
When the proposed coder is used in a packet network, the robustness is directly related to PLR and SR. Figures 7-9 show some reconstructed Lena and Peppers images at different values of SR and PLR. It can be observed that most of the visual information of the original images can be recovered even when $SR=0.2$ and $PLR=0.3$. This demonstrates that the proposed coder possesses high robustness against packet loss. Besides, the coder does not result in blocking artifacts. In the aforementioned experiments, the packet size is set to 100 while the block size of SM is 32$\times$32. In fact, the robustness is more or less related to both values.
As analyzed previously, there are three factors causing the PSNR difference in exploring the relationship between SR and PLR. Yet these factors arise from the packet size $m$. Intuitively, with an increasing $m$, the PSNR value descends to some extent. This conjecture is justified by Fig. 10, where the parameter settings are $SR=0.6$ and $PLR=0.3$. The smaller the packet size, i.e., the more the number of descriptions, the better the reconstructed image quality is. Naturally, the best case is that each measurement forms a description. When the packet size is between 0 and $3 \times {10^4}$, the PSNR value drops with the reduction in the number of packets. However, when the packet size is larger than $3 \times {10^4}$, the PSNR virtually has no change. This is because that the number of packets is basically reduced to two and remains unchanged. If one of these two packets is lost, it means that half of the successive measurements are sampled. This successional sampling violates the randomness of the down-sampling operator ${\bf{D}}$. The analyses indicate that if the transmission channel allows a small quantity of descriptions and the PLR is too large, for instance, only two descriptions and $PLR \ge 0.3$, the proposed coder cannot be regarded as an efficient multiple description coder. In order to fix this problem, Charlie has to improve the SR. Consider an extreme scenario that $SR=1$, i.e., full redundancy without compression, the encoding process is changed to ${\bf{y = F}}{{\bf{x}}_{en}}$. Such an encoder cannot be guaranteed by the theory of SRM and a great many successive measurements' loss will substantially affect the quality of the reconstructed image. Fortunately, a solution has been developed to cope with this scenario. Associating a realization of down-sampling operator ${\bf{D}}$ that truncates the first or $M$ randomly-selected elements after arbitrarily permuting the signal, Charlie introduces a new random permutation ${\bf{R'}}$ known by Bob. The present encoding form is ${\bf{y = R'F}}{{\bf{x}}_{en}}$. When a packet containing many successive measurements is lost, Bob receives the information ${\bf{\hat y = }}\beta {\bf{R'F}}{{\bf{x}}_{en}}$. Let ${\bf{D'}} = \beta {\bf{R'}}$, which can be considered as a down-sampling operator, then ${\bf{\hat y = D'F}}{{\bf{x}}_{en}}$. In other words, the PLR is the very SR. Even if $PLR = 0.8$, which is equivalent to $SR = 0.8$, the reconstructed image quality is still visually acceptable.
The purpose of having the measurement matrix in a block mode is to reduce storage space and computational complexity at the cost of a lower quality of the recovered signal. In the proposed coder, we investigate PSNR versus the block size of SM when $SR=0.6$ and $PLR=0.05$, as shown in Fig. 11. The greater the block size, the higher the PSNR is. However, the rate of increase is quite slow. Meanwhile, a larger block size of SM needs more memory and consumes more resources. Consequently, a trade-off between them is required. In general, the block size of SM is set as 32$\sim$256.
\section{Conclusion}
A novel and robust coder for processing encrypted images against packet loss has been designed. It is different from the existing approaches of the robust coding of natural images and the compression of encrypted images. The proposed coder based on SRM is composed of three parts: permutation-based encryption by Alice, encoding with structural matrix by Charlie, and joint decryption and decoding by Bob. In addition, we have investigated the relationship between the proposed and the existing methods. Two other cryptographic applications of SRM have also been suggested. In the performance evaluation, we have explored the relationship between packet loss rate and sampling rate. A feasible approach for quantizing the compressive measurements of encrypted images has been introduced. Finally, we have investigated the robustness of the proposed coder at different parameter settings. It has been verified that our coder can be considered as an efficient multiple description coder with a number of descriptions to resist packet loss.
\bibliographystyle{IEEEtr}
|
\section{Introduction}
Hybrid type-logical grammars \cite{kl12gap,kl13emp,kl13coord} are a relatively new framework in
computational linguistics, which combines insights from the Lambek
calculus \cite{lambek} and lambda grammars
\cite{oehrle,muskens01lfg,muskens03lambda} --- lambda grammars are also called, depending on the authors,
abstract categorial grammars \cite{groote01acg} and linear grammars \cite{pollard11linear}, though with somewhat
different notational conventions\footnote{I prefer the term lambda
grammars, since I think it most clearly describes the
system. Though the term abstract categorial grammars appears to be
more common, and I use it from time to time in this article, I will argue in Section~\ref{inadeq} that abstract
categorial grammars/lambda grammars are \emph{unlike} all other
versions of categorial grammars in important ways.}. The resulting
combined system solves some know problems of both the Lambek calculus
and of lambda grammars and the additional expressiveness of hybrid
type-logical grammars permits the treatment of linguistic phenomena
such as gapping
which have no satisfactory solution in either subsystem.
The goal of this paper is to prove that hybrid
type-logical grammars are a fragment of first-order linear logic. This
embedding result has several important consequences: it not only
provides a simple new proof theory for the calculus, thereby clarifying
the proof-theoretic foundations of hybrid type-logical grammars, but, since the
translation is simple and direct, it also provides several new parsing
strategies for hybrid type-logical grammars. Second, NP-completeness of
hybrid type-logical grammars follows immediately.
The main embedding result also sheds new light on problems with
lambda grammars, which are a subsystem of hybrid
type-logical grammars and hence a special case of the translation into
first-order linear logic.
Abstract categorial grammars are attractive both because of
their simplicity --- they use the simply typed lambda calculus, one of the most widely used
tools in formal semantics, to compute surface structure (strings)
as well as to compute logical form (meanings) --- and because of the fact that they
provide a natural account of quantifier scope and extraction; for
both, the analysis is superior to the Lambek calculus analysis. So it
is easy to get the impression that lambda grammars are an unequivocal improvement
over Lambek grammars.
In reality, the picture is much more nuanced: while lambda grammars
have some often discussed advantages over Lambek grammars, there are
several cases --- notably coordination, but we will see in Section~\ref{inadeq}
that this is true for any analysis where the Lambek calculus uses
non-atomic arguments --- where the Lambek grammar analysis
is clearly superior. Many key examples illustrating the elegance
of categorial grammars with respect to the syntax-semantics interface fail to have a satisfactory treatment in abstract
categorial grammars.
However, whether or not lambda grammars are an improvement over the
Lambek calculus is ultimately not the most important question. Since
there is a large number of formal systems which improve upon the
Lambek calculus, it makes much more sense to compare lambda grammars
to these extensions, which include, among many others, Hybrid
Type-Logical Grammars, the Displacement calculus \cite{mvf11displacement} and multimodal
type-logical grammars \cite{mmli,M95}. These extended Lambek calculi all keep the things
that worked in the Lambek calculus but improve on the analysis in ways
which allow the treatment of more complex phenomena in syntax and
especially in the syntax-semantics interface. Compared to these systems, the inadequacies of lambda
grammars are evident: even for the things lambda grammars do right (quantifier scope and
extraction), there are phenomena, such as reflexives and gapping, which
are handled by the same mechanisms as quantifier scope and extraction
in alternative theories, yet which cannot be adequately handled by
lambda grammars. The abstract categorial grammar treatment suffers
from problems of overgeneration and problems at the syntax-semantics
interface unlike any other categorial grammar. I will discuss some
possible solutions for lambda grammars, but it is clear that a major
redesign of the theory is necessary. The most painless solution seems
to be a move either to hybrid type-logical grammars or directly to
first-order linear logic: both are simple, conservative extensions
which solve the many problems of lambda grammars while staying close
to the spirit of lambda grammars.
This paper is structured as follows. Section~\ref{sec:mill1} will
introduce first-order linear logic and Section~\ref{sec:stl} will
provide some background about the simply typed lambda
calculus. These two introductory sections can be skimmed by people familiar
with first-order linear logic and the simply typed lambda calculus respectively. Section~\ref{sec:hybrid} will introduce hybrid type-logical
grammars and in Section~\ref{sec:equiv} we will give a translation of
hybrid type-logical grammars into first-order linear logic and prove
its correctness. Section~\ref{sec:comparison} will then compare
the Lambek calculus and several of its extensions through their
translations in first-order linear logic. This comparison points to a
number of potential problems for lambda grammars. We will
discuss these problems, as well as some potential solutions in Section~\ref{inadeq}.
Finally, the last section will contain some concluding remarks.
\section{First-order Linear Logic}
\label{sec:mill1}
Linear logic was introduced by \citeasnoun{Girard} as a logic
which restricts the structural rules which apply freely in classical logic.
The multiplicative, intuitionistic fragment of first-order linear
logic (which in the following, I will call either MILL1 or simply
first-order linear logic), can be seen as a resource-conscious version of first-order
intuitionistic logic. Linear implication, written $A \multimap B$, is
a variant of intuitionistic implication $A\Rightarrow B$ with the
additional constraint that the $A$ argument formula is used
\emph{exactly once}. So, looking at linear logic from the context of grammatical analysis, we would assign an intransitive verb the
formula $np\multimap s$, indicating it is a formula which
combines with a single $np$ (noun phrase) to form an $s$ (a sentence).
Linear logic is a commutative logic. In the context of language
modelling, this means our languages are closed under permutations of
the input string, which does not make for a good linguistic principle
(at least not a good \emph{universal} one and a principle which is at
least debatable even in languages which allow relatively free word order). We need some way to restrict or
control commutativity. The Lambek calculus \cite{lambek} has the simplest such restriction: we drop
the structural rule of commutativity altogether. This means linear implication
$A \multimap B$ splits into two implications: $A\backslash B$, which
looks for an $A$ to its left to form a $B$, and $B/A$, which looks for
an $A$ to its right to form a $B$. In the Lambek calculus, we would
therefore refine the assignment to
intransitive verbs from $np\multimap s$ to $np\backslash s$, indicating the intransitive verb
is looking for the subject to its left.
In first-order linear logic, we can choose a more versatile solution,
namely using first-order variables to encode word order. We assign
atomic formulas a pair of string positions: $np$ becomes $np(0,1)$,
meaning it is a noun phrase spanning position 0 (its leftmost
position) to 1 (its rightmost position). Using pairs of (integer)
variables to represent strings is standard in parsing algorithms.
The addition of quantifiers makes things more interesting. For example, we can assign the formula $\forall x. n(3,x)
\multimap np(2,x)$ to
a determiner ``the'' which spans positions $2,3$. This means it is looking for a noun which starts
at its right (that is the leftmost position of this noun is the
rightmost position of the determiner, 3) but ends at any position $x$
to produce a noun phrase which starts at position 2 (the leftmost
position of the determiner) and ends at position $x$ (the rightmost
position of the noun). Combined with a noun $n(3,4)$, this would
allow us to instantiate $x$ to 4 and produce $np(2,4)$. In other
words, the formula given to the determiner indicates it is looking for
a noun to its right in order to produce a noun phrase, using a form of
``concatenation by instantiation of variables'' which should be familiar to
anyone who has done some logic programming or who has a basic
familiarity with parsing in general \cite{PS87,dedpar}. Similarly, we can assign an
intransitive verb at position 1,2 the formula $\forall y. np(y,1)
\multimap s(y,2)$ to indicate it is looking for a noun phrase to its
left to form a sentence, as the Lambek calculus formula $np\backslash
s$ for intransitive verbs does --- this correspondence between
first-order linear logic and the Lambek calculus is fully general and
discussed fully in \cite{mill1} and briefly in the next section.
\subsection{MILL1}
\label{sec:nounif}
After this informal introduction to first-order linear logic, it is
time to be a bit more precise. We will not need function symbols in
the current paper, so
\emph{terms} are either variables denoted $x, y, z, \ldots $ (a
countably infinite number) or constants, for which I
will normally use integers $0, 1, \ldots$, giving an $m$-word string
$m+1$ string positions, from $0$ to $m$. The atomic formulas are of the form $a(t_1,
\ldots, t_m)$ with $t_i$ terms, $a$ a predicate symbol (we only need a
finite, typically smal, number of predicate symbols, often only the
following four: $n$ for noun,
$np$ for noun phrase, $s$ for sentence, $pp$ for predicate phrase) and $m$ its
arity. Our language does \emph{not} contain the identity relation symbol ``=''. Given this set of atomic formulas $\mathcal{A}$ and the set of variables $\mathcal{V}$, the set of formulas is
defined as follows\footnote{We need neither the multiplicative
conjunction $\otimes$ nor the existential quantifier $\exists$ in
this paper, though adding them to the logic poses no problems. The
natural deduction rules for $\exists$ and $\otimes$ are slightly
more complicated than those for $\forall$ and $\multimap$ but the
basic proof net building blocks don't change, see
for example \cite{quant,mill1,moot13lambek}.}.
$$
\mathcal{F} ::= \mathcal{A} \; | \; \mathcal{F} \multimap \mathcal{V} \; | \; \forall \mathcal{V}. \mathcal{F}
$$
We treat formulas as syntactically equivalent up to renaming of bound
variables, so substituting $\forall y. A [x:=y]$ (where $A$ does not
contain $y$ before this substitution is made) for $\forall x. A$
inside a formula $B$ will produce an equivalent formula, for
example $\forall x. a(x) \equiv \forall y. a(y)$.
Table~\ref{tab:millnd} shows the natural deduction rules for
first-order linear logic. The variable $x$ in the $\forall E$ and
$\forall I$ rules is called the \emph{eigenvariable} of the
rule. The $\forall I$ rule has the condition that the variable $y$
which is replaced by the eigenvariable
does not occur in undischarged hypotheses of the proof and that $x$
does not occur in $A$ before the substitution is made\footnote{It is
sometimes more convenient to use the following $\forall I$
rule $$\infer[\forall I^*]{\forall x. A}{A}
$$ \noindent with the condition there are no free occurrence of $x$
in open hypotheses. The rule of Table~\ref{tab:millnd} is more
convient in the following section when we use meta-variables, where
it becomes ``replace all occurrences of a (meta-)variable by $x$,
then quantify over $x$''.}.
Throughout this paper, we will use the standard convention in
first-order (linear) logic \cite{quant,empires,bpt} that every occurrence of
a quantifier $\forall$, $\exists$ in a sequent uses
a distinct variable and in addition that no variable occurs both free
and bound in a sequent.
\begin{table}
$$
\begin{array}{ccc}
\infer[\multimap E]{B}{A & A \multimap B} &&
\infer[\multimap I]{A \multimap B}{\infer*{B}{[A]^i}} \\
\\
\infer[\forall E]{A[x:=t]}{\forall x. A} & &
\infer[\forall I^*]{\forall x. A}{A[y:=x]} \\
\end{array}
$$
\caption{Natural deduction rules for first-order linear logic}
\label{tab:millnd}
\end{table}
As shown in \cite{mill1}, we can translate Lambek calculus sequents
and formulas into first-order linear logic as follows.
\begin{gather*}
A_1, \ldots, A_n \vdash B = \\
\| A_1 \|^{0,1}, \ldots \| A_n \|^{n-1,n} \vdash \| B \|^{0,n}
\end{gather*}
\begin{align*}
\| a \|^{x,y} & = a(x,y) \\
\| A/ B \|^{x,y} &= \forall z. \| B \|^{y,z} \multimap \| A \|^{x,z} \\
\| B\backslash A \|^{y,z} &= \forall x \| B \|^{x,y} \multimap \| A \|^{x,z}
\end{align*}
The integers 0 to $n$ represent the positions of the formulas in the
sequent and the translations for complex formulas introduce
universally quantified variables. The translation for $A/B$ states
that if we have a formula $A/B$ at positions $x,y$ then for any $z$ if
we find a formula $B$ at positions $y,z$ (that is, to the immediate
right of our $A/B$ formula) then we have an $A$ at positions $x,z$,
starting at the left position of the $A/B$ formula and ending at the
right position of the $B$ argument. In
other words, a formula $A/B$ is something which combines with a $B$
to its right to form an $A$, just like its Lambek calculus counterpart.
Using this translation, we can see that the first-order linear logic
formulas used for the determiner and the intransitive verb in the
previous section correspond to the translations of $np/n$ at position
$2,3$ and $np\backslash s$ at position $1,2$ respectively.
To give a simple example of a first-order linear logic proof, we shown
a derivation of ``every student ran'', corresponding to the Lambek
calculus sequent.
$$
(s/(np\backslash s))/n, n, np\backslash s \vdash s
$$
We first translate the sequent into first-order linear logic.
$$
\| (s/(np\backslash s))/n \|^{0,1}, \| n\|^{1,2}, \| np\backslash s
\|^{2,3} \vdash \| s \|^{0,3}
$$
Then translate the formulas as follows.
$$
\forall y. [ n(1,y) \multimap \forall z. [ \forall x. [np(x,y) \multimap
s(x,z)] \multimap s(0,z) ] ], n(1,2), \forall v. [ np(v,2) \multimap
s(v,3) \vdash s(0,3) ]
$$
We can then show that ``every student ran'' is a grammatical sentence
under these formula assignments as follows.
$$
\infer[\multimap E]{s(0,3)}{\infer[\forall E]{\forall x. [np(x,2) \multimap
s(x,3)] \multimap s(0,3)}{
\infer[\multimap E]{\forall z. [ \forall x. [np(x,2) \multimap
s(x,z)] \multimap s(0,z) ]}{
\infer[\forall E]{n(1,2) \multimap \forall z. [ \forall x. [np(x,2) \multimap
s(x,z)] \multimap s(0,z) ]}{\forall y. [ n(1,y) \multimap \forall z. [ \forall x. [np(x,y) \multimap
s(x,z)] \multimap s(0,z) ] ]} & n(1,2)}} & \forall v. [np(v,2)
\multimap s(v,3)]}
$$
The application of the final $\multimap E$ rule is valid, since $\forall x. [np(x,2) \multimap
s(x,3)] \equiv \forall v. [np(v,2)
\multimap s(v,3)]$.
\begin{definition}[Universal closure]\label{def:cl} If $A$ is a formula we denote the set of free
variables of $A$ by $\textit{FV}(A)$.
For an antecedent
$\Gamma=A_1,\ldots,A_n$, $\textit{FV}(\Gamma) = \textit{FV}(A_1)
\cup \dots \cup \textit{FV}(A_n)$.
The \emph{universal closure} of a formula $A$ with $\textit{FV}(A) = \{x_1,\ldots,x_n\}$, denoted
$\textit{Cl}(A)$, is the formula $\forall x_1 \ldots \forall x_n . A$.
The \emph{universal closure} of a formula $A$ \emph{modulo antecedent}
$\Gamma$, written $\textit{Cl}_{\Gamma}(A)$, is defined by universally
quantifying over the free variables in $A$ which do not occur in
$\Gamma$. If $\textit{FV}(A) \setminus \textit{FV}(\Gamma) = \{
x_1,\ldots, x_n\}$, then $\textit{Cl}_{\Gamma}(A) = \forall x_1
\ldots \forall x_n . A$.
\end{definition}
\begin{proposition}\label{prop:cl} $\Gamma \vdash A$ iff $\Gamma \vdash \textit{Cl}_{\Gamma}(A)$.
\end{proposition}
\paragraph{Proof} If the closure modulo $\Gamma$ prefixes $n$ universal
quantifiers to $A$, we can go from $\Gamma \vdash A$ to $\Gamma \vdash
\textit{Cl}_{\Gamma}(A)$ by using the $\forall I$ rule $n$ times (the
quantified variables added for the closure have been chosen to respect the condition on the rule) and
in the opposite direction by using the $\forall E$ rule $n$ times.
\hfill \ensuremath{\Box}
\subsection{MILL1 with focusing and unification}
\label{sec:unif}
The $\forall E$ rule, as formulated in the previous section, has the
disadvantage that it requires us to choose a term $t$ with which to
replace $x$ and that making the right choice for $t$ requires some
insight into how the resulting formula will be used in the rest of the proof. In
the example of the preceding section we need to make two such
``educated guesses'': we instantiate $y$ to 2 to allow the
elimination rule with minor premiss $n(1,2)$ and we instantiate $z$ to 3
to produce the desired conclusion $s(0,3)$.
The
standard solution to automate this process in first-order logic
theorem proving is to change the $\forall E$ rule:
instead of directly replacing the quantified variable by the ``right''
choice, we replace it by a meta-variable (I
will use the Prolog-like notation $A$, $B$, $\ldots$ for these
variables, or, when confusion with the notation $A$ and $B$ for arbitrary
formulas is possible $C$, $D$, $E$, $\ldots$, $V$, $W$, $X$,
$\ldots$).
These meta-variables will represent our current
knowledge about the term with which we will replace a given quantified
variable. The MGU we compute for the endsequent will correspond to
the most general instantiations of these variables in the given proof (that is, all other
instantiations can be obtained from this final MGU by means of additional substitutions).
The $\multimap E$ rule \emph{unifies} the $B$ formulas of the
argument and minor premiss of the rule (so the two occurrences of $B$
need only be unifiable instead of identical). Remember that the unification
of two atomic formulas $a(x_1,\ldots,x_m)$ and $b(y_1,\ldots,y_n)$ is
only defined when $a=b$ and $m=n$ and that unification tries to find
the most general instantiation of all free variables such that $x_i =
y_i$ (for all $1 \leq i \leq n = m$) and fails if no such
instantiation exists. The presence of an explicit quantifier presents a
complication, but only a minor one: bound variables are treated just
like constants which, in addition, must respect the variable condition.
More precisely, the unification of
two formulas is defined as
follows.
\begin{align*}
\textit{unify}(a(x_1,\ldots,x_n), a(y_1,\ldots,y_n)) &=
\textit{unify}(x_i,y_i)\ \textrm{for all}\ 1 \leq i \leq n \\
\textit{unify}(A_1\multimap B_1, A_2\multimap B_2) &= \textit{unify}(A_2,
A_1),\textit{unify}(B_1, B_2)\\
\textit{unify}(\forall x.A, \forall y.B) & = \textit{unify}(A, B[y:=x])
\end{align*}
The $\forall$ case assumes there are no free occurrences of $x$ in $B$
before substitution. It is defined in such a way that it is
independent of the actual variable names used for the quantifier (as
mentioned, we use a different variable for each occurrence of a
quantifier) and bound occurrences of $x_i$ and $y_i$ are treated as
constants in the $\textit{unify}(x_i,y_i)$ clause, subject to the
following condition:
if we compute a substitution $D:=x$ for a
formula $A$ and $x$ is not free
for $D$ in $A$ then unification \emph{fails}. In other words, the substitution cannot introduce new
bound variables, so for example $\forall y. a(D,y)$ and $\forall
z. a(z,z)$ fail to unify, since $D$ is not free for $y$ in $\forall
y. a(D,y)$, and therefore we cannot legally substitute $y$ for $D$
since it would result in an ``accidental capture'', creating a new
bound occurrence of $y$.\footnote{In such cases, substitution
succeeds but does nothing and subsequent unification fails, since
the formulas are not alphabetic variants after substitution.}
As second problem with natural deduction proof search is that we can
have subproofs like the following.
$$
\begin{array}{cc}
\infer[\forall E]{a(y)}{\infer[\forall I]{\forall x. a(x)}{a(y)}} &
\infer[\multimap E]{B}{\infer[\multimap I_i]{A\multimap B}{\infer*{B}{[A]^i}} & A}
\end{array}
$$
In both cases, we introduce a connective and then immediately
eliminate it. A natural deduction proof is called \emph{normal} if is
does not contain any subproof of the forms shown above. One of the classic results for natural deduction is
\emph{normalization} which states that we can eliminate such detours
\cite{glt,bpt}. In the case of linear logic, removing such detours is
even guaranteed to decrease the size of the proof.
We use a form of
\emph{focalized} natural deduction \cite{focus,fnd}, which is a
syntactic variant of natural deduction guaranteed to generate only
normal natural deduction proofs. We use two turnstiles, the
negative $\vdash_n$ and the positive $\vdash_p$ (for the reader
familiar with focused
proofs, $\Gamma \vdash C \Downarrow$ corresponds to $\Gamma \vdash_n
C$ and $\Gamma \vdash C \Uparrow$ to $\Gamma \vdash_p C$).
We will call a sequent $\Gamma \vdash_p C$ a \emph{positive} sequent
(and $C$ a positive formula) and a sequent $\Gamma \vdash_n C$ a
\emph{negative} sequent (and $C$ a negative formula).
\editout{
using the
following translations for negative and positive sequents. $\| A_1
\|^- ,\ldots \| A_n \|^- \vdash_n \| B \|^-$ and $\| A_1 \|^- ,\ldots
\| A_n \|^- \vdash_p \| B \|^+$, with the positive and negative
formula translations defined as
shown below.
\begin{align*}
\| p \|^- &= p \\
\| \forall x. A \|^- &= \| A [x:= D] \|^- \\
\| A \multimap B \|^- &= \| A \|^+ \multimap \| B \|^-
\end{align*}
\begin{align*}
\| p \|^+ &= p \\
\| \forall x. A \|^+ &= \forall x. \| A \|^+ \\
\| A \multimap B \|^+ &= \| A \|^- \multimap \| B \|^+
\end{align*}
$$
\infer[\multimap I]{\forall y . np(2,y) \multimap \forall x. [ np(x,1) \multimap
s(x,y) \vdash np(2,B) \multimap np(A,1) \multimap s(A,B)}{
\infer[\forall E]{np(2,B),\forall y . np(2,y) \multimap \forall x. [ np(x,1) \multimap
s(x,y) \vdash np(A,1) \multimap s(A,B) }{
\infer[\multimap E]{np(2,B),\forall y . np(2,y) \multimap \forall x. [ np(x,1) \multimap
s(x,y) \vdash \forall v. np(v,1) \multimap s(v,B) }{np(2,B) \vdash np(2,B) &
\infer[\forall E]{\forall y . np(2,y) \multimap \forall x. [ np(x,1) \multimap
s(x,y) ] \vdash np(2,B) \multimap \forall v . [np(v,1) \multimap s(v,B)]}{\forall y . np(2,y) \multimap \forall x. [ np(x,1) \multimap
s(x,y) ] \vdash \forall w . np(2,w) \multimap \forall v. [ np(v,1) \multimap s(v,w) ]}}}}
$$
}
\editout{
For any formula $A$, the translation of an axiom $\| A \|^- \vdash \| A \|^+$,
allows us to uniquely recover the explicitly quantified formula $A$, since all
quantifiers appear explicitly either in $\| A \|^+$ or in $\| A
\|^-$. In addition, it is easy to verify that for all formulas $A$, $\| A \|^- \vdash A$ and
that $A\vdash \| A\|^+$ (and hence $\| A \|^- \vdash \| A \|^+$): we need only check the positive universal
quantifier and show that given $A[x:=D] \vdash \| A' \|^+$ we can derive
$A[x:=D]\vdash \forall x. \| A' \|^+$, which is a simple application of the
$\forall I$ rule since the left hand side of the turnstile does not
contain $x$, only occurrences of $D$.
}
\begin{table}
\begin{center}
\textbf{Lexicon}
\end{center}
\vspace{-1\baselineskip}
$$
A \vdash_n A
$$
\vspace{.3\baselineskip}
\begin{center}
\textbf{Axiom/Hypothesis}
\end{center}
\vspace{-1\baselineskip}
$$
A \vdash_n A
$$
\vspace{.3\baselineskip}
\begin{center}
\textbf{Shift Focus}
\end{center}
\vspace{-1\baselineskip}
$$
\infer[\pm]{\Gamma \vdash_p A}{\Gamma \vdash_n A}
$$
\vspace{.3\baselineskip}
\begin{center}
\textbf{Logical Rules}
\end{center}
\vspace{-.5\baselineskip}
$$
\infer[\multimap E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta) \vdash_n \mathbb{s}(A)}{\Gamma \vdash_n B \multimap
A & \Delta \vdash_p B}
$$
$$
\infer[\multimap I]{\Gamma \vdash_p B \multimap A}{\Gamma, B \vdash_p A}
$$
$$
\infer[\forall E]{\Gamma \vdash_n A[x:=D]}{\Gamma\vdash_n \forall
x. A}
$$
$$
\infer[\forall I^*]{\Gamma \vdash_p \forall x.A}{\Gamma \vdash_p A[y:=x]}
$$
\caption{Focused first-order linear logic with unification}
\label{tab:rulesimpl}
\end{table}
Table~\ref{tab:rulesimpl} shows the rules of first-order linear logic
in this format. For the lexicon rule, we require that the formula $A$ is closed. The formula $A$ of the hypothesis rule \emph{can} contain free variables.\editout{
hypothesis rule contains no constants
but can contain free
variables, though free variables are constrained to occur once on each side of
the turnstile\footnote{That is, we cannot have a hypothesis like
$a(x,x)\vdash a(x,x)$. As we will see, this restriction corresponds
to a type of linearity of the variables and, though non-standard, it
will simplify the proofs which follow. Note that this restriction is
compatible with the standard analysis of \textit{wh}-movement in
categorial grammars: an $np$ trace will start as $np(x,y)\vdash
np(x,y)$ and get further instantiated by unification later in the
proof.}.}
For the $\forall I$ rule, $y$ is either a variable or a meta-variable
which has no free occurrences in any undischarged hypothesis.
For the $\multimap E$ rule,
$\mathbb{s}$ is the most general unifier of $\langle \Gamma, B\rangle$ and
$\langle \Delta, B\rangle$. That is, we unify the two occurrences of $B$ in
their respective contexts, using unification for complex formulas as
defined above.
The resulting most general unifier is then applied
to the two contexts and to $A$ (replacing, if necessary, any variables shared
between $A$ and $B$ in the formula $A$).
We can see from the rules that axioms start negative and stay negative
as long as they are the major premiss of a $\multimap E$ rule or the
premiss of a $\forall E$ rule. We must
switch to positive sequents to use the introduction rules or to use
the sequent as the minor premiss of a $\multimap E$
rule.
The ``detour'' subproofs we have seen
above cannot receive
a consistent labeling: the formula $A\multimap B$ is the
conclusion of a $\multimap I$ rule and must therefore be on the
right-hand side
of a positive sequent, however, it is also the major premiss of a
$\multimap E$ rule and must therefore be on the right-hand side of a negative
sequent (it is easily verified there is no way to transform a positive
sequent into a negative sequent, however the point is that the
\emph{original} detour receives an inconsistent labeling)
$$
\begin{array}{cc}
\infer[\forall E]{\vdash_n a(y)}{\infer[???]{\vdash_n \forall x. a(x)}{\infer[\forall I]{\vdash_p \forall x. a(x)}{\vdash_p a(y)}}} &
\infer[\multimap E]{\vdash_n B}{\infer[???]{\vdash_n A \multimap
B}{\infer[\multimap I_i]{\vdash_p A\multimap B}{\infer*{\vdash_p
B}{[\vdash_n A]^i}}} & \vdash_p A}
\end{array}
$$
\begin{definition}\label{def:track}
A \emph{principal branch} is a sequence of negative sequents which starts at a
hypothesis, then follows all elimination rules from (major) premiss
to conclusion ending at a focus shift rule (this corresponds to the
normal notion of principal branch from e.g.\ \cite{glt}; a sequence of
negative sequents can only pass through the major premiss of a
$\multimap E$ rule and through the single premiss of a $\forall E$ rule).
A \emph{track} is a path
of negative sequents followed by a focus shift followed by a path of
positive sequents. A track ends either
in the conclusion of the proof or in the minor premiss of a
$\multimap E$ rule.
The \emph{main track} of a proof is the track which ends in its conclusion
(these definitions corresponds to the
standard notion of track and main track in normal proofs, see e.g.\ \cite{bpt}).
\end{definition}
This suggests a relation between focused proofs and normal natural
deduction proofs, which is made explicit in the following two propositions.
\editout{
But first, we prove a small lemma to show that the translation of
first-order linear logic formulas defined above is well-behaved with
respect to the focus switch rule of Table~\ref{tab:rulesimpl}.}
\editout{
\begin{lemma}\label{lem:shift} For every first-order linear logic formula $A$ and
context $\Gamma$, the derived rule
$$\infer[pn]{\| \Gamma \|^- \vdash_p \| C \|^+}{\| \Gamma \|^- \vdash_n
\| C \|^-}$$
\noindent is valid.
\end{lemma}
\paragraph{Proof} By induction on the structure of $C$.
If $C$ is an atomic formula, then the rule reduces to the focus
switch rule, since the positive and negative translation produce the
same formula.
If $C\equiv A\multimap B$, then by induction hypothesis, the statement
holds for both $A$ and $B$, and we can derive the desired rule as
follows (IH denotes application of the induction hypothesis).
$$
\infer[=_{\textit{def}}]{\| \Gamma \|^- \vdash_p \| A \multimap B
\|^+}{\infer[\multimap I]{\| \Gamma
\|^- \vdash_p \| A \|^- \multimap \| B \|^+}{\infer[IH]{\| \Gamma
\|^- , \| A \|^- \vdash_p \| B \|^+}{\infer[\multimap E]{\| \Gamma
\|^- , \| A \|^- \vdash_n \| B \|^-}{\infer{\| \Gamma \|^-
\vdash_n \| A \|^+ \multimap \| B \|^-}{\| \Gamma\|^- \vdash_n
\| A \multimap B \|^-} & \infer[IH]{\| A
\|^- \vdash_p \| A \|^+ }{\infer{\| A \|^- \vdash_n \| A \|^-}{}}}}}}
$$
If $C\equiv \forall x .A$, then we proceed as follows.
$$
\infer[=_{\textit{def}}]{\| \Gamma \|^- \vdash_p \| \forall x. A
\|^+}{\infer[\forall I]{\| \Gamma \|^- \vdash_p \forall x. \| A
\|^+}{\infer[IH]{\| \Gamma \|^- \vdash_p \| A[x:=D]
\|^+}{\infer[=_{\textit{def}}]{\| \Gamma\|^- \vdash_n \|
A[x:=D] \|^-}{\| \Gamma \|^- \vdash_n \| \forall x. A\|^-}}}}
$$
Since $D$ is guaranteed not to occur outside of $A$ in the proof, the
application of $\forall I$ is valid.
\hfill \ensuremath{\Box}}
\begin{proposition}\label{prop:focus} For every natural deduction proof of $\Gamma\vdash B$, there is a
focused natural deduction proof with unification of $\Gamma \vdash_p B$.
\end{proposition}
\paragraph{Proof} We first transform the natural deduction
proof of $\Gamma\vdash B$ into a normal natural deduction proof, then
proceed by induction on the length of the proof and show that we can
create both a proof of $\Gamma \vdash_p B$ and a substitution
$\mathbb{s}$. We proceed by induction on the depth of the proof.
If $d=1$, we have an axiom or hypothesis rule, which we translate as
follows.
$$
\infer[\pm]{A\vdash_p A}{A\vdash_n A}
$$
If $d > 1$ we proceed by case analysis on the last rule.
The only case which requires some attention is the $\multimap E$
case. Given that the proof is normal, we have a normal (sub)proof
which ends in a $\multimap E$ rule. We are therefore on the principal
branch of this subproof and we know that a principal branch starts with
an axiom/lexicon rule then passes only $\forall E$ rules and $\multimap E$ rules through
their major premiss. Hence, the last rule producing the major premiss
in the original proof must either have been an axiom/lexicon rule or
an elimination rule for $\multimap$ or $\forall$.
Now induction hypothesis gives us a proof $\delta_1$ of $\Gamma
\vdash_p B\multimap A$ and a proof $\delta_2$ of $\Delta \vdash_p
B$. However, given that the last rule of the proof which produces
$\delta_1$ was either axiom/lexicon, the $\forall E$ rule or the
$\multimap E$ rule --- all of which have negative sequents as their
conclusion --- the last rule of
$\delta_1$ must have been the focus shift rule. Removing this focus
shift rule produces a valid proof $\delta_{1'}$ of $\Gamma \vdash_n B\multimap A$,
which we can combine with the proof $\delta_2$ of $\Delta\vdash_p B$ as follows.
$$
\infer[\pm]{\Gamma, \Delta \vdash_p A}{\infer{\Gamma, \Delta
\vdash_n A}{\infer*[\delta_{1'}]{\Gamma\vdash_n B \multimap A}{} & \infer*[\delta_2]{\Delta \vdash_p B}{}}}
$$
Note that this is again a proof which ends with a focus shift rule.
Since the original proof uses the stricter notion of \emph{identity}
(instead of unifiability) for the $B$ formulas, we need not change the
substitution we have computed so far and therefore leave $\Gamma$,
$\Delta$ and $A$ unchanged.
For the $\forall E$ rule, induction hypothesis gives us a proof
$\delta$ of $\Gamma \vdash_p \forall x. A$, by reasoning similar to
the case for $\multimap E$, we know the last rule of $\delta$ was a
focus shift rule, which we can remove, then extend the proof as follows.
$$
\infer[\pm]{\Gamma\vdash_p A[x:=D]}{\infer[\forall
E]{\Gamma\vdash_n A[x:=D]}{\infer*[\delta]{\Gamma \vdash_n \forall x. A}{}}}
$$
Adding the substitution $D := t$ (where $t$ is the term used for the
in the original $\forall E$ rule) to the unifier.
The cases for $\forall I$ and $\multimap I$ are trivial, since we can
extend the proof with the same rule.
\hfill \ensuremath{\Box}
\begin{proposition} For every focused natural deduction proof, there
is a natural deduction proof.
\end{proposition}
\paragraph{Proof} If we remove the focus shift rule and replace both
$\vdash_n$ and $\vdash_p$ by $\vdash$ then we only need to give
specific instantiations for the $\forall E$ rules. The most general
unifier $\mathbb{s}$ computed for the complete proof gives us such values for each
(negatively) quantified variable (if wanted, remaining meta-variables can be
replaced by free variables). \hfill \ensuremath{\Box}
The following is a standard property of normal natural deduction
proofs (and therefore of focused natural deduction proofs).
\begin{proposition} Focused proofs satisfy the \emph{subformula}
property. That is, any formula occurring in a proof of $\Gamma
\vdash_p B$ (or $\Gamma \vdash_n B$) is a subformula either of
$\Gamma$ or of $B$.
\end{proposition}
The following proposition is easily verified by induction on $A$ and
using the correspondence between natural deduction proofs and $\lambda$-terms.
\begin{proposition} We can restrict the focus shift rule to atomic
formulas $A$. When we do so, we only produce long normal form proofs
(which correspond to beta normal eta long
lambda terms).
\end{proposition}
\editout{
\begin{proposition}\label{propa} We can assume, without loss of generality, that a
conclusion of the form $\forall x. (B \multimap A)$ has been obtained
by a consecutive use of the $\multimap I$ and $\forall I$ rules.
\end{proposition}
\paragraph{Proof} This is immediate from the setup of the focalised
system, given that we can restrict the use of the focus shift rule to
atomic formulas. So we are in the following situation
$$
\infer[\forall I]{\Gamma \vdash_p \forall x. (B\multimap A)}{\Gamma
\vdash_p B \multimap A}
$$
\noindent and, since $B\multimap A$ is a complex formula, we cannot
change to a negative sequent. Hence, the only available rule given the
polarity of the sequent is
$\multimap I$.
An easy alternative proof can be given using the proof nets of the
following section, which is the basic observation that combinations of
positive links can be treated as a unit (ie.\ as a single, combined rule). \hfill \ensuremath{\Box}
}
The proof from the previous section looks as follows in the
unification-based version of first-order linear logic, though we use a
form with implicit antecedents to economize on horizontal space and
to make comparison with the proof of the previous section
easier. This proof produces the most general unifier $Y=2$, $Z=3$,
corresponding to the explicit instantiations for $y$ and $z$ at the $\forall E$ rules
in the previous proof.
\medskip
\noindent\scalebox{.83}{$$
\infer[\pm]{\vdash_p s(0,3)}{\infer[\multimap E]{\vdash_n s(0,3)}{
\infer[\forall E]{\vdash_n \forall x. [np(x,2)\multimap s(x,Z)] \multimap s(0,Z)}{\infer[\multimap E]{\vdash_n \forall z. [ \forall x. [np(x,2) \multimap
s(x,z)] \multimap s(0,z) ]}{
\infer[\forall E]{\vdash_n n(1,Y) \multimap \forall z. [ \forall x. [np(x,Y) \multimap
s(x,z)] \multimap s(0,z) ]}{\vdash_n \forall y. [ n(1,y) \multimap
\forall z. [ \forall x. [np(x,y) \multimap
s(x,z)] \multimap s(0,z)] ] } & \infer[\pm]{\vdash_p n(1,2)}{\vdash_n
n(1,2)}}} & \infer[\pm]{\vdash_p \forall v. [np(v,2)
\multimap s(v,3)]}{\vdash_p \forall v. [np(v,2)
\multimap s(v,3)] }}}
$$}
\medskip
\editout{
\medskip
\noindent\scalebox{.83}{$$
\infer[\multimap E]{s(0,3)}{
\infer[\multimap E]{\forall x. [np(x,2) \multimap
s(x,Z)] \multimap s(0,Z) }{
\infer[\pm]{n(1,Y) \multimap \forall x. [np(x,Y) \multimap
s(x,Z)] \multimap s(0,Z) }{\forall y. [ n(1,y) \multimap \forall z. [ \forall x. [np(x,y) \multimap
s(x,z)] \multimap s(0,z) ] ]} & n(1,2)} & \infer[\forall I]{\forall x. [np(x,2)
\multimap s(x,3)]}{\infer[\pm]{np(V,2)\multimap s(V,3) }{ \forall v. [ np(v,2)\multimap s(v,3) ]}}}
$$}
\medskip
}
Restricting
focus shift ($\pm$) to atomic formulas, produces the following
proof in long normal form.
Remark that our hypothesis in this proof
is not $np(V,2)$ but $np(U,W)$ which unifies with $np(V,2)$ at the
$\multimap E$ rule immediately below it.
\noindent\scalebox{.75}{$$
\infer[\pm]{\vdash_p s(0,3)}{
\infer[\multimap E]{\vdash_n s(0,3)}{
\infer[\forall E]{\vdash_n \forall x. [np(x,2) \multimap
s(x,Z) ] \multimap s(0,Z)}{
\infer[\multimap E]{ \vdash_n \forall z. [\forall x. [np(x,2) \multimap s(x,z)] \multimap s(0,z) ]}{
\infer[\forall E]{\vdash_n n(1,Y) \multimap \forall z. [ \forall x. [np(x,Y) \multimap s(x,z)] \multimap s(0,z)] }{\vdash_n
\forall y. [ n(1,y) \multimap \forall z. [ \forall x. [np(x,y) \multimap
s(x,z)] \multimap s(0,z) ] ]}
& \infer[\pm]{\vdash_p n(1,2)}{\vdash_n n(1,2)}
}} &
\!\!\!\!\!\!\!\!\!\!\!\!\infer[\forall I]{\vdash_p \forall w. [np(w,2)\multimap s(w,3)]}{
\infer[\multimap I_1]{\vdash_p np(V,2)\multimap s(V,3)}{
\infer[\pm]{\vdash_p s(V,3)}{\infer[\multimap E]{\vdash_n s(V,3)}{
\infer[\pm]{\vdash_p np(U,W)}{\infer[\textit{Hyp}_1]{\vdash_n np(U,W)}{}}
& \infer[\forall E]{\vdash_n np(V,2)\multimap
s(V,3)}{\vdash_n \forall v. np(v,2) \multimap s(v,3)}
}}
}
}
}}
$$}
\subsection{Proof Nets}
\label{sec:pn}
Proof nets are an elegant alternative to natural deduction and an
important research topic in their own right; for reasons of space we
provide only an informal introduction --- the reader interested in more
detail is referred to \cite{llintro} for an
introduction and to \cite{multiplicatives,empires} for detailed proofs in the context of
linear logic and to \cite{pnlambek,diss,mr12lcg} for introductions in
the context of categorial grammars and the Lambek calculus. Though
proof nets shine especially for the $\exists$ and $\otimes$ rules (where the
natural deduction formulation requires commutative conversions to
decide proof equivalence), they are a useful alternative in the
$\forall$ and $\multimap$ case as well since they provide an easy
combinatorial way to do proof search and therefore make arguments
about non-derivability of statements and serve to count the number of readings.
\citeasnoun{quant} shows that the proof nets of multiplicative linear
logic \cite{Girard,multiplicatives} have a simple extension to the
first-order case. Essentially, a proof net is a graph labeled with (polarized
occurrences of) the
(sub)formulas of a sequent $\Gamma \vdash C$, subject to some conditions we will
discuss below. Obviously, not all graphs labeled with formulas correspond to
derivable statements. However, we can characterize the proof nets
among the larger class of proof structures (graphs labeled with
formulas which, contrary to proof nets, do \emph{not} necessarily correspond
to proofs) by means of simple graph-theoretic properties.
The basic building blocks of proof structures are \emph{links}, as
shown in
Figure~\ref{fig:links}. We will call the formulas displayed below the link their
\emph{conclusion} and the formulas displayed above it their
premisses. The axiom link (top left) has no premisses and two
conclusions, the cut link has no conclusions and two premisses, the
binary logical links have two premisses ($A$ and $B$) and one
conclusion $A\multimap B$ and the unary logical links have one premiss
$A$ and one conclusion $\forall x.A$. We will call $x$ the
eigenvariable of the link and require that all links use distinct
variables
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (forallnc) {$\overset{-}{\forall x. A}$};
\node (forallnp) [above=2em of forallnc] {$\overset{-}{A[x:=t]}$};
\draw (forallnc) -- (forallnp);
\node (forallpc) [right=7em of forallnc] {$\overset{+}{\forall x. A}$};
\node (forallpp) [above=2em of forallpc] {$\overset{+}{A_{\rule{0pt}{1.55ex}}}$};
\draw [dotted] (forallpc) -- (forallpp);
\node (lollinc) [above=7em of forallnc] {$\overset{-}{A\multimap B}$};
\node (tmplnl) [left=0.66em of lollinc] {};
\node (alollin) [above=2.5em of tmplnl] {$\overset{+}{A}$};
\draw (lollinc) -- (alollin);
\node (tmplnr) [right=0.66em of lollinc] {};
\node (blollin) [above=2.5em of tmplnr] {$\overset{-}{B}$};
\draw (lollinc) -- (blollin);
\node (lollipc) [above=7em of forallpc] {$\overset{+}{A\multimap B}$};
\node (tmplpl) [left=0.66em of lollipc] {};
\node (alollip) [above=2.5em of tmplpl] {$\overset{-}{A}$};
\node (tmplpr) [right=0.66em of lollipc] {};
\node (blollip) [above=2.5em of tmplpr] {$\overset{+}{B}$};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path [clip] (lollipc.center) circle (2.5ex) [reverseclip];
\end{pgfinterruptboundingbox}
\draw [dotted] (lollipc.center) -- (blollip);
\draw [dotted] (lollipc.center) -- (alollip);
\end{scope}
\begin{scope}
\path [clip] (alollip) -- (lollipc.center) -- (blollip);
\draw (lollipc.center) circle (2.5ex);
\end{scope}
\node (anx) [above=3.5em of alollin] {$\overset{-}{A}$};
\node (anp) [above=3.5em of blollin] {$\overset{+}{A}$};
\node (anxa) [above=1em of anx] {};
\node (anpa) [above=1em of anp] {};
\draw (anx) -- (anxa.center) -- (anpa.center) -- (anp);
\node (cnx) [above=3.5em of alollip] {$\overset{-}{A}$};
\node (cnp) [above=3.5em of blollip] {$\overset{+}{A}$};
\node (cnxb) [below=1em of cnx] {};
\node (cnpb) [below=1em of cnp] {};
\draw (cnx) -- (cnxb.center) -- (cnpb.center) -- (cnp);
\end{tikzpicture}
\end{center}
\caption{Links for proof structures in the $\forall$, $\multimap$
fragment of first-order linear logic.}
\label{fig:links}
\end{figure}
Given a statement $A_1, \ldots, A_n \vdash C$
we can unfold the formulas using the logical links of the figure, using the
negative links for the $A_i$ and the positive link for $C$. Since
there is only one type of link for each combination of
connective/polarity, we unfold our formulas
deterministically\footnote{For the negative $\forall$ this is not
immediately obvious, since we need to choose a suitable term $t$. We
will discuss this case below but we will essentially use
meta-variables and
unification just like we did for natural deduction in Section~\ref{sec:unif}.}, until
we end up at the atomic formulas and have produced a ``formula
forest'', a sequence of formula decomposition trees labeled with some additional
information (polarity labels and dashed lines), which is sometimes called a \emph{proof frame}.
We turn this proof
frame into a \emph{proof structure} by connecting atomic formulas of
opposite polarity in such a way there is a perfect matching between
the positive and negative atoms. This step can already fail, for
example if the number of positive and negative occurrences of an
atomic formula differ but also because of incompatible atomic
formulas
like $a(0,1)$ and $a(x,1)$, with $x$ the eigenvariable of a
$\forall^+$ link. More generally, it can be the case that there is no
coherent substitution which allows us to perform a complete matching
of the atomic formulas using axiom links. These restrictions on the
instantiations of variables are a powerful tool for proof search \cite{moot07filter,moot13lambek}.
Proof structures are essentially
graphs where some of the links are drawn with dashed lines; the binary
dashed lines are paired, as indicated by the connecting arc. We will
call the dashed logical links ($\forall^+$ and $\multimap^+$) the
\emph{positive} links and the solid logical links ($\forall^-$ and $\multimap^-$)
the \emph{negative} links. The terms positive and negative
links only apply to the logical links; the axiom and cut link
are neither positive nor negative. A proof structure containing only
negative logical links is just a graph labeled with polarized formulas.
Figure~\ref{fig:psex} shows the proof net which corresponds to the
natural deduction proof of Section~\ref{sec:nounif}. To save space, we
have noted only the main connective at each link; the full formula can
be obtained unambiguously from the context. We have also been free
in the way we ordered the premisses of the $\multimap$ links, which
allows us to give a planar presentation of the axiom links, much like
Lambek calculus proof nets. However, there is no planarity requirement
in the proof net calculus; the first-order variables offer more
flexibility than simple planarity. For the $\forall^-$ links, we have
annotated the substitutions next to the link. If we use a
unification-based presentation, as we did for natural deduction in
Section~\ref{sec:unif}, we can ``read off'' these substitutions from the most
general unifier computed for the axioms (as opposed to natural
deduction, the axioms and not the $\multimap E$ rule, which
corresponds to the $\multimap^-$ link, are responsible for the
unification of variables).
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (b1) at (6,0) {$\overset{-}{\forall y}$};
\node (bv1) at (6,1.2) {$\overset{-}{\multimap}$};
\node (npos) at (7,2.4) {$\overset{+}{n(1,2)}$};
\node (fn1) at (5,2.4) {$\overset{-}{\forall z}$};
\node (fi1) at (5,3.6) {$\overset{-}{\multimap}$};
\node (s1) at (4,4.8) {$\overset{-}{s(0,3)}$};
\node (fx) at (6,4.8) {$\overset{+}{\forall x}$};
\node (pi) at (6,6) {$\overset{+}{\multimap}$};
\node (s2) at (5,7.2) {$\overset{+}{s(x,3)}$};
\node (np1) at (7,7.2) {$\overset{-}{np(x,2)}$};
\node (n2) at (9,0) {$\overset{-}{n(1,2)}$};
\node (b3) at (12,0) {$\overset{-}{\forall v}$};
\node (f3) at (12,1.2) {$\overset{-}{\multimap}$};
\node (np2) at (11,2.4) {$\overset{+}{np(x,2)}$};
\node (s3) at (13,2.4) {$\overset{-}{s(x,3)}$};
\node (g) at (14,0) {$\overset{+}{s(0,3)}$};
\draw (b1) -- (bv1);
\node (y) at (6.6,0.6) {$y:=2$};
\draw (bv1) -- (npos);
\draw (bv1) -- (fn1);
\draw (fn1) -- (fi1);
\node (z) at (5.6,3.0) {$z:=3$};
\draw (fi1) -- (s1);
\draw (fi1) -- (fx);
\draw[dashed] (fx) -- (pi);
\draw[dashed] (pi) -- (s2);
\draw[dashed] (pi) -- (np1);
\draw (b3) -- (f3);
\node (v) at (12.6,0.6) {$v:=x$};
\draw (f3) -- (np2);
\draw (f3) -- (s3);
\draw (npos) -- (7,3.2) -- (9,3.2) -- (n2);
\draw (np1) -- (7,8.0) -- (11,8.0) -- (np2);
\draw (s2) -- (5,8.6) -- (13,8.6) -- (s3);
\draw (s1) -- (4,9.2) -- (14,9.2) -- (g);
\end{tikzpicture}
\end{center}
\caption{Proof net corresponding to the natural deduction proof of
Section~\ref{sec:nounif}}
\label{fig:psex}
\end{figure}
A proof structure is a \emph{proof net} if the statement $A_1,\ldots, A_n
\vdash C$ is derivable, that is, given the proof of
Section~\ref{sec:nounif}, we know the proof structure of
Figure~\ref{fig:psex} is a proof net. However, this definition is
not very useful, since it depends on finding a proof in some other
proof system; we would like to use the proof structure itself
to directly decide whether or not the statement is derivable. However, it is possible to distinguish the proof
nets from the other proof structures by simple graph-theoretic
properties. To do so, we first introduce some auxiliary notions, which
turn the graph-like proof structures into standard graphs. Since
axiom, cut and the
negative links already produce normal graphs ($\multimap^-$
corresponds to two edges, all other links to a single edge in the graph), we only need a way to
remove the positive links.
\begin{definition}
A \emph{switching} is a choice for each positive link as follows.
\begin{itemize}
\item For each $\multimap^+$ link, we choose one its
premisses ($A$ or $B$).
\item For each $\forall^+$ link, we choose \emph{either} its premiss $A$ \emph{or} any of
the formulas in the proof structure containing a free occurrence of
the eigenvariable of the link.
\end{itemize}
A given a switching $s$, a \emph{correction graph} is a proof structure
where we replace all dashed links by a link from the conclusion of the
link to the formula chosen by the switching $s$.
\end{definition}
\begin{theorem} \cite{quant} A proof structure is a \emph{proof net} iff all its
correction graphs are acyclic and connected.
\end{theorem}
Defined like this, it would seem that deciding whether or not a proof
structure is a proof net is rather complicated: there are potentially
many correction graphs --- we have two independent possibilities for each $\multimap^+$
link and generally at least two subformulas containing the
eigenvariable of each $\forall^+$ link, giving $2^n$ correction graphs
for $n$ positive links --- and we need verify all of them. Fortunately,
there are very efficient alternatives: linear time in the
quantifier-free case \cite{murong,pnlinear} and at most squared time, though possibly better,
in the case with quantifiers \cite{moot13lambek}.
Going back to the example shown in Figure~\ref{fig:psex}, we can see
that there are two positive links and twelve correction graphs: there
are six free occurrences of $x$ --- four in atomic formulas and two
additional occurrences in the
conclusions ($\multimap^+$ and $\multimap^-$) which combine these
atomic formulas into $np(x,2) \multimap s(x,2)$ --- times the two independent possibilities for switching
$\multimap^+$ left or right. We can verify that all twelve
possibilities produce acyclic, connected graphs. Removing the positive
links splits the graph into three connected components: the single
node labeled $\multimap^+$ (representing $(np(x,2) \multimap
s(x,2))^+$), a component containing the intransitive verb ending at
the axioms to $s(x,3)^+$ and $np(x,2)^-$ and a final component
containing the rest of the graph, ending at the conclusion of the
$\forall^+$ link (which has been disconnected from its premiss). Now, any
switching for the $\multimap^+$ link will connect its isolated
conclusion node
to the component containing $s(x,3)^+$ and $np(x,2)^-$ (via one or
the other of these nodes), leaving two connected components. Finally,
all free occurrences of the variable $x$ occur in this newly created
component, therefore any choice for a switching of the $\forall^+$
link will join these disconnected components into a single, connected component.
Since each choice connected two disjoint components, we have not
generated any cycles.
We can also show that this is the only possible proof structure for
the given logical statement: there is only one choice for the $n$
formulas, one choice for the $np$ formulas though two choices for the
$s$ formulas. However, the alternative proof structure would link $s(0,z)$ to
$s(x,z)$ (for some value of $z$), which fails because $x$, being the
eigenvariable of a $\forall^+$ link, cannot be instantiated to 0.
As a second example, let's show how we can use correction graphs to
show underivability.
Though it is clear that the switching for the universal quantifier
must refer to free occurrences of its eigenvariable somewhere (as do
its counterparts in natural deduction and sequent calculus), it is not
so easy to find a small example in the $\forall,\multimap$ fragment
where this condition is necessary to show underivability, since
finding a global instantiation of the variables is already a powerful
constraint on proof structures. However, the existential quantifier
and the universal quantifier differ only in the labeling of
formulas for the links and we need the formula labeling only for
determining the free variables.
A proof structure of the underivable sequent $(\forall
x. a(x)) \multimap b \nvdash \exists y. [ a(y) \multimap b ] $ is
shown in Figure~\ref{fig:noproof}. It is easy to verify this is the
unique proof structure corresponding to this sequent. This sequent is used for
computing the prenex normal form of a formula in classical logic
(replacing $\multimap$ by $\Rightarrow$), but it is invalid in
intuitionistic logic and linear logic since it depends on the
structural rule of right contraction.
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (b1) at (1.5cm,0cm) {$\overset{-}{\forall x. a(x) \multimap b}$};
\node (lax) at (0cm,1.5cm) {$\overset{+}{\forall x. a(x)}$};
\node (lb) at (3cm,1.5cm) {$\overset{-}{b}$};
\node (g) at (6cm,0cm) {$\overset{+}{\exists y. [ a(y) \multimap b ] }$};
\node (g2) at (6cm,1.5cm) {$\overset{+}{a(x) \multimap b}$};
\draw (g) -- (g2);
\node (ga) at (7.5cm,3cm) {$\overset{-}{a(x)}$};
\node (gb) at (4.5cm,3cm) {$\overset{+}{b}$};
\node (la) at (0cm,3cm) {$\overset{+}{a(x)}$};
\draw (ga) -- (7.5cm,4.5cm) -- (0cm,4.5cm) -- (la);
\draw (gb) -- (4.5cm,4cm) -- (3cm,4cm) -- (lb);
\draw[dashed] (lax) -- (la);
\draw (b1) -- (lax);
\draw (b1) -- (lb);
\draw[dashed] (g2) -- (ga);
\draw[dashed] (g2) -- (gb);
\end{tikzpicture}
\end{center}
\caption{Proof structure which is not a proof net}
\label{fig:noproof}
\end{figure}
In order to show the sequent is invalid in linear logic, it suffices
to find a switching such that the corresponding correction graph
either contains a cycle or is disconnected. Figure~\ref{fig:cycle} shows a
correction graph for the proof structure of Figure~\ref{fig:noproof}
which is both cyclic and disconnected: the axiom $a(x) \vdash a(x)$
is not connected to the rest of the structure and the connection
between $\forall x. a(x)$ and $a(x) \multimap b$ produces a cycle,
since there is a second path to these two formulas through the axiom
$b\vdash b$.
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (b1) at (1.5cm,0cm) {$\overset{-}{\forall x. a(x) \multimap b}$};
\node (lax) at (0cm,1.5cm) {$\overset{+}{\forall x. a(x)}$};
\node (lb) at (3cm,1.5cm) {$\overset{-}{b}$};
\node (g) at (6cm,0cm) {$\overset{+}{\exists y. [ a(y) \multimap b ] }$};
\node (g2) at (6cm,1.5cm) {$\overset{+}{a(x) \multimap b}$};
\draw (g) -- (g2);
\node (ga) at (7.5cm,3cm) {$\overset{-}{a(x)}$};
\node (gb) at (4.5cm,3cm) {$\overset{+}{b}$};
\node (la) at (0cm,3cm) {$\overset{+}{a(x)}$};
\draw (ga) -- (7.5cm,4.5cm) -- (0cm,4.5cm) -- (la);
\draw (gb) -- (4.5cm,4cm) -- (3cm,4cm) -- (lb);
\draw (b1) -- (lax);
\draw (b1) -- (lb);
\draw (g2) -- (gb);
\draw plot [smooth] coordinates {(lax.north) (2.5cm,4.0cm) (5.0cm,4.0cm) (g2.north)};
\end{tikzpicture}
\end{center}
\caption{A cyclic and disconnected correction graph for the proof
structure of Figure~\ref{fig:noproof}}
\label{fig:cycle}
\end{figure}
This concludes our brief introduction to proof nets for first-order
linear logic. We refer the reader to Appendix~A of \cite{glt} for
discussion about the relation between proof nets and natural deduction.
\section{Basic Properties of the Simply Typed Lambda Calculus}
\label{sec:stl}
Before introducing hybrid type-logical grammars, we will first review
some basic properties of the simply typed lambda calculus which will
prove useful in what follows. This section is not intended as a
general introduction to the simply typed lambda calculus: we will
assume the reader has at least some basic knowledge such as can be
found in Chapter~3 of \cite{glt} or other textbooks and some knowledge
about substitution and most general unifiers. For more detail, and for
proofs of the lemmas and propositions of this section, the reader is
referred to \cite{hindley}.
A remark on notation: we will use $\rightarrow$ exclusively as a type
constructor (also when we know we are using it to type a linear lambda term) and
$\multimap$ exclusively as a logical connective.
\begin{definition} A lambda term $M$ is a \emph{linear} lambda term
iff
\begin{enumerate}
\item for every subterm $\lambda x. N$ of $M$, $x$ has exactly one
occurrence in $N$ (in other words, each abstraction binds exactly one variable occurrence),
\item all free variables of $M$ occur exactly once.
\end{enumerate}
\end{definition}
Table~\ref{tab:typing} lists the Curry-style typing rules for the
linear lambda calculus. For the $\rightarrow E$ rule, $\Gamma$ and $\Delta$
cannot share term variables; for the $\rightarrow I$ rule, $\Gamma$ cannot contain
$x$ (ie.\ $\Gamma, x^{\alpha}$ must be a valid context).
\begin{table}
$$
\begin{array}{c}
x^{\alpha} \vdash x:\alpha \\
\\
\infer[\rightarrow E]{\Gamma,\Delta\vdash (M\,N):\beta}{\Gamma \vdash
M:\alpha\rightarrow \beta & \Delta \vdash N:\alpha}\\
\\
\infer[\rightarrow I]{\Gamma\vdash \lambda x.M:\alpha\rightarrow\beta}{\Gamma,x^{\alpha} \vdash M:\beta}
\end{array}
$$
\caption{Curry-style typing rules for the linear lambda calculus}
\label{tab:typing}
\end{table}
\begin{proposition}\label{prop:linear} For linear lambda terms, we have the following:
\begin{enumerate}
\item\label{prop:fv} When $M$ is a linear lambda term and $\Gamma \vdash
M:\alpha$ a deduction of $M$, then the variables occurring in
$\Gamma$ are exactly the free variables of $M$.
\item If $M$, $N$ are linear lambda terms which do not share free
variables then $(M\,N)$ is a linear lambda term.
\item If $M$ is a linear lambda term with a free occurrence of $x$
then $\lambda x.M$ is a linear lambda term.
\item If $M$ is a linear lambda term and
$M\twoheadrightarrow_{\beta\eta} N$ then $N$ is a linear lambda
term.
\end{enumerate}
\end{proposition}
\editout{
Inversion Lemma, case (iii). Proposition 1.2.3, case (iii) from
Barendregt e.a.
\begin{proposition} If $\Gamma\vdash \lambda x.M:A$, then $A \equiv
B\rightarrow C$ and $\Gamma, x:B \vdash M:C$.
\end{proposition}
}
\begin{lemma}[Substitution]\label{lem:subst} If $\Gamma, x:\alpha \vdash M:\beta$,
$\Delta\vdash N:\alpha$ and $\Gamma$ and $\Delta$ are compatible
(ie.\ there are no conflicting variable assignments and therefore
$\Gamma,\Delta$ is a valid context),
then $\Gamma,\Delta \vdash M[x:=N]:\beta$.
\end{lemma}
The following two results are rather standard, we can find them in
\cite{hindley} as Lemmas~2C1 and 2C2.
\begin{lemma}[Subject Reduction]\label{lemma:sr} Let $M \twoheadrightarrow_{\beta\eta}
N$, then $\Gamma \vdash M:\alpha \Rightarrow \Gamma \vdash N:\alpha$
\end{lemma}
\begin{lemma}[Subject Expansion]\label{lemma:se} Let $M \twoheadrightarrow_{\beta\eta}
N$ with $M$ a linear lambda term, then $\Gamma \vdash N:\alpha \Rightarrow \Gamma \vdash M:\alpha$
\end{lemma}
\subsection{Principal types}
\label{sec:pt}
The main notions from Chapter~3 of \cite{hindley} are the following.
\begin{definition}[Principal type] A \emph{principal type} of a term
$M$ is a type $\alpha$ such that
\begin{enumerate}
\item for some context $\Gamma$ we have
$\Gamma \vdash M:\alpha$
\item if $\Gamma' \vdash M:\beta$, then there is a substitution
$\mathbb{s}$ such that $\mathbb{s}(\alpha) = \beta$.
\end{enumerate}
\end{definition}
\begin{definition}[Principal pair] A \emph{principal pair} for a term
$M$ is a pair $\langle \Gamma, \alpha\rangle$ such that $\Gamma
\vdash M:\alpha$ and for all $\beta$ such that $\Gamma \vdash
M:\beta$ there is a substitution $\mathbb{s}$ with $\mathbb{s}(\alpha) = \beta$
\end{definition}
\begin{definition}[Principal deduction] A \emph{principal deduction}
for a term $M$ is a derivation $\delta$ of a statement $\Gamma
\vdash M:\alpha$ such that every other derivation with term $M$ is
an instance of $\delta$ (ie.\ obtained by globally applying a
substitution $\mathbb{s}$ to all types in the proof).
\end{definition}
From the definitions above, it is clear that if $\delta$ is a
principal deduction for $\Gamma \vdash M:\alpha$ then $\Gamma,\alpha$
is a principal pair and $\alpha$ a principal type of $M$.
If $M$ contains
free variables $x_1,\ldots x_n$ we can compute the principal
type $\alpha_1 \rightarrow \ldots (\alpha_n \rightarrow \beta)$ of the closed term $\vdash \lambda
x_1,\ldots x_n.M$ which is the same as the principal type for
$x_1^{\alpha_1},\ldots,x_n^{\alpha_n}\vdash M^{\beta}$.
\subsection{The principal type algorithm}
The principal type algorithm of \citeasnoun{hindley} is defined as
follows. It is slightly more general and computes principal
\emph{deductions}. It takes as input a lambda term $M$ and outputs either its
principal type $\alpha$ or fails in case $M$ is untypable. We closely
follow Hindley's presentation, keeping his numbering but restricting
ourselves to linear lambda terms; we omit his
correctness proof of the algorithm.
We proceed by induction on the construction of $M$.
{
\renewcommand{\theenumi}{\Roman{enumi}}
\renewcommand{\theenumii}{\alph{enumii}}
\renewcommand{\theenumiii}{\arabic{enumiii}}
\begin{enumerate}
\item If $M$ is a variable, say $x$, then we take an unused type variable
$\alpha$ and return $x^{\alpha} \vdash x:\alpha$ as principal
deduction.
\item If $M$ is of the form $\lambda x. N$ and $x$ occurs in $N$ then
we look at the principal deduction $\delta$ of $N$ by induction hypothesis
: if we fail to compute a principal deduction for $N$ then there is
no principal deduction for $\lambda x. N$ either. If such a
deduction $\delta$ does exist, then we can extend it as follows.
$$
\infer[\multimap I]{\Gamma \vdash \lambda
x.N:\alpha\rightarrow\beta}{\infer*[\delta]{x^{\alpha},\Gamma\vdash N:\beta}{}}
$$
\item $M$ is of the form $\lambda x. N$ and $x$ does not occur in $N$;
this case cannot occur since it violates the condition on linear
lambda terms (we must bind exactly one occurrence of $x$ in $N$), so we fail.
\item\label{pt:appl} $M$ is of the form $(N\, P)$. If the algorithm fails for either
$N$ or $P$, then $M$ is untypable and we fail. If not, induction
hypothesis gives us a principal proof $\delta_1$ for $\Gamma \vdash
N:\gamma'$ and a principal proof $\delta_2$ for $\Delta\vdash
P:\gamma$. If necessary, we rename type variables if $\Gamma$ and $\Delta$ such
that $\Gamma$ and $\Delta$ have no type variables in common. Since
$M$ is linear, $N$ and $P$ cannot share term variables.
\begin{enumerate}
\item If $\gamma'$ is of the form $\alpha \rightarrow \beta$
then we compute the most general unifier $\mathbb{s}$ of $\langle \Gamma,
\alpha\rangle$ and $\langle \Delta,\gamma\rangle$. If this fails the
term is untypable; if not we combine the proofs as
follows.
$$
\infer[\rightarrow E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta) \vdash (N\,
P):\mathbb{s}(\beta)}{\infer*[\mathbb{s}(\delta_1)]{\mathbb{s}(\Gamma)\vdash N:\mathbb{s}(\alpha)\rightarrow\mathbb{s}(\beta)}{}
& \infer*[\mathbb{s}(\delta_2)]{\mathbb{s}(\Delta)\vdash P:\mathbb{s}(\gamma)}{}}
$$
\item If $\gamma'$ is a type variable, then we compute the most
general unifier $\mathbb{s}$ of $\langle \Gamma,\gamma'\rangle$ and $\langle
\Delta,\gamma\rightarrow\beta\rangle$ (with $\beta$ a fresh type
variable). If this succeeds and the term is typable, we can produce
its principal proof as follows.
$$
\infer[\rightarrow E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta) \vdash (N\,
P):\mathbb{s}(\beta)}{\infer*[\mathbb{s}(\delta_1)]{\mathbb{s}(\Gamma)\vdash N:\mathbb{s}(\gamma)\rightarrow\mathbb{s}(\beta)}{}
& \infer*[\mathbb{s}(\delta_2)]{\mathbb{s}(\Delta)\vdash P:\mathbb{s}(\gamma)}{}}
$$
\end{enumerate}
\end{enumerate}
}
The main utility of principal types in the current paper
is given by the coherence theorem.
\begin{theorem}[Coherence]\label{thm:coherence} Suppose $\Gamma \vdash N:\alpha$ and let
$\alpha$ be a principal type of $N$ then $\forall
P \, \forall \Gamma' \subseteq \Gamma\ \ \Gamma' \vdash P:\alpha
\ \Longrightarrow\ P \equiv_{\beta\eta} N$
\end{theorem}
The coherence theorem states that a principal type determines a
lambda term uniquely (up to $\beta\eta$ equivalence). Since we work in
a linear system, where weakening is not allowed, we only need the
special case $\Gamma' = \Gamma$. This special case of
Theorem~\ref{thm:coherence} is the following: if $\Gamma\vdash N:\alpha$ with $\alpha$ a principal type of
$N$ then for any $P$ such that $\Gamma \vdash P:\alpha$ we have that
$P \equiv_{\beta\eta} N$.
In brief, the principal type algorithm allows us to
compute the principal type of a given typable lambda term, whereas the
coherence theorem allows us to reconstruct a lambda term (up to
$\beta\eta$ equivalence) from a principal type.
\begin{definition} We say a sequent $\Gamma\vdash C$ is \emph{balanced} if all
atomic types occurring in the sequent occur exactly twice.
\end{definition}
The following lemmas are easy consequences of 1) the Curry-Howard
isomorphism between linear lambda terms and Intuitionistic Linear
Logic (ILL), which allows us to interpret the linear type constructor
``$\rightarrow$'' as the logical connective ``$\multimap$'' 2) the correspondence between (normal) natural deduction proofs
and (cut-free) proof
nets and 3) the fact that renaming the conclusions of the axiom links
in a proof net gives another proof net.
\begin{lemma}\label{lemma:balanced} If $M$ is a linear
lambda term with free variables $x_1,\ldots,x_n$ then the principal
type $\alpha_1 \rightarrow \ldots (\alpha_1 \rightarrow \beta)$ of $\lambda x_1 \ldots
\lambda x_n.M$ is balanced. Hence the principal type of
$x_1^{\alpha_1},\ldots x_n^{\alpha_n} \vdash M^{\beta}$ is balanced.
\end{lemma}
\paragraph{Proof} Compute the natural deduction proof of $M$ and
convert it to a ILL proof net. By subject reduction
(Lemma~\ref{lemma:sr}), normalization/cut elimination keeps the type
$\alpha$ invariant. Let $P$ be the cut-free proof net which corresponds to
the natural deduction proof of $M$ and which has the same type as $M$.
We obtain a balanced proof net by using a different atomic formula for
all axiom links. From this proof net, we can obtain all other types of
$M$ by renaming the axiom links (allowing for non-atomic axiom links), hence it is a principal type and it
is balanced by construction. \hfill \ensuremath{\Box}
\begin{lemma}If $M$ is a beta-normal lambda term with free variables
$x_1,\ldots,x_n$ and if $\lambda x_1,\ldots,\lambda x_n.M$ has a
balanced typing then $M$ is linear.
\end{lemma}
\paragraph{Proof} If $\lambda x_1,\ldots,\lambda x_n.M$ has a
balanced typing, then from this typing we can construct a unique cut-free ILL
proof net of $\lambda x_1,\ldots,\lambda x_n.M$. Since it is an ILL
proof net, this lambda term must be linear and therefore $M$ as well.
\hfill \ensuremath{\Box}
\subsection{Examples}
\label{sec:exc}
To illustrate the principal type algorithm, we give two examples in
this section.
As a first example, we compute the principal proof of $\textbf{C}
\equiv \lambda f . \lambda x .
\lambda y . ((f\, y)\, x)$
as follows.
$$
\infer[\rightarrow I]{\vdash \lambda f. \lambda x.\lambda y. ((f\,y)\,x):(\beta\rightarrow\alpha\rightarrow\gamma)\rightarrow\alpha\rightarrow\beta\rightarrow\gamma}{
\infer[\rightarrow I]{f^{\beta\rightarrow\alpha\rightarrow\gamma}\vdash \lambda x.\lambda y. ((f\,y)\,x):\alpha\rightarrow\beta\rightarrow\gamma}{
\infer[\rightarrow I]{f^{\beta\rightarrow\alpha\rightarrow\gamma},x^{\alpha}\vdash\lambda y. ((f\,y)\,x):\beta\rightarrow\gamma}{
\infer[\rightarrow E]{f^{\beta\rightarrow\alpha\rightarrow\gamma},y^{\beta},x^{\alpha}\vdash
((f\,y)\,x):\gamma}{\infer[\rightarrow
E]{f^{\beta\rightarrow\gamma_1},y^{\beta}\vdash
(f\,y):\gamma_1}{f^{\gamma_0} \vdash f:\gamma_0 & y^{\beta} \vdash
y:\beta} & x^{\alpha}\vdash x:\alpha}}}}
$$
The substitutions $\gamma_0 := \beta\rightarrow\gamma_1$ (for the topmost $\rightarrow
E$ rule) and $\gamma_1 :=
\alpha\rightarrow \gamma$ (for the bottom $\rightarrow E$ rule) have been left implicit
in the proof.
As a second example, the principal proof of $l^{2\rightarrow 1} \vdash \lambda
O. \lambda S . \lambda z. (S\,(l\, (O\, z)))$ is the following.
$$
\infer[\rightarrow I]{l^{2\rightarrow 1}\vdash \lambda O. \lambda S. \lambda
z. (S\,(l\,(O\, z))):(\beta\rightarrow 2)\rightarrow(1\rightarrow\alpha)\rightarrow\beta\rightarrow\alpha}{
\infer[\rightarrow I]{l^{2\rightarrow 1},O^{\beta\rightarrow
2}\vdash \lambda S. \lambda z. (S\,(l\,(O\, z))):(1\rightarrow\alpha)\rightarrow\beta\rightarrow\alpha}{
\infer[\rightarrow I]{S^{1\rightarrow\alpha},l^{2\rightarrow 1},O^{\beta\rightarrow
2}\vdash \lambda z. (S\,(l\,(O\, z))):\beta\rightarrow\alpha}{
\infer[\rightarrow E]{S^{1\rightarrow\alpha},l^{2\rightarrow 1},O^{\beta\rightarrow
2},z^{\beta}\vdash (S\,(l\,(O\, z))):\alpha}{S^{\alpha_2} \vdash S:\alpha_2 &
\infer[\rightarrow E]{l^{2\rightarrow 1},O^{\beta\rightarrow 2},z^{\beta}\vdash (l\,(O\, z)):1}{l^{2\rightarrow 1}\vdash l:2\rightarrow 1 &\infer[\rightarrow E]{O^{\beta\rightarrow\alpha_1},z^{\beta} \vdash (O\, z):\alpha_1}{O^{\alpha_0}\vdash O:\alpha_0 & z^{\beta}\vdash z:\beta}}}}}}
$$
The substitutions $\alpha_0 := \beta\rightarrow\alpha_1$, $\alpha_1 := 2$,
$\alpha_2 := 1\rightarrow\alpha$ (of the three $\rightarrow E$ rules, from top to
bottom) have again been left implicit.
\editout{
\section{Correctness}
Given a mapping $m$ from the arguments of atomic predicates to atomic subtypes of a type (ie. statements of the form $\textit{inf}(A,B,C,D) \equiv (C \rightarrow B) \rightarrow D \rightarrow A$ for all atomic formulas).
Linear grammar (lexicalized) to MILL1 grammar: compute the principal type of all lexical entries, match type and formula, translating $\rightarrow$ to $\multimap$ and atomic types according to the mapping $m$. Translate the goal of type $n \rightarrow 0$ (for some integer $n > 0$) to $s(0,n)$.
MILL1 grammar proof (which is a translation of a linear grammar, and
therefore only contains $\multimap$, negative $\forall$ and atomic
formulas in the range of $m$) to linear grammar proof: translate the formulas (without quantifiers) to types, since each variable/constant occurs at most once negatively and at most once positively the coherence theorem guarantees that there is a unique $\lambda$-term in $\beta$-normal $\eta$-long form which corresponds to this type.
In fact, we have that the sequent $R\rightarrow L \vdash \alpha$, where $\alpha$ is type corresponding to the input formula and $L$ and $R$ are the left and right position constants is a sequent where each atomic formula has \emph{exactly} one negative and \emph{exactly} one positive occurrence and therefore a unique proof net, which corresponds to a $\lambda$-term (if this unique proof structure is \emph{not} a proof net, the type is uninhabited).
QUESTION: are there conditions on the formula which guarantee the type
is inhabited? The fact that we can apply the coherence theorem here
already presupposes that each variable has exactly one positive and
exactly one negative occurrence (according to the images in $m$) and
that the type has one negative occurrence of $L$ and one positive
occurrence of $R$. $R\rightarrow L \vdash \alpha$ being a tautology of
(implicational) linear logic would suffice (in addition to the
condition on the occurrences of variables), excluding things like
$5\rightarrow 4 \vdash (B \rightarrow B) \rightarrow 5 \rightarrow 4$, but \emph{not} $5\rightarrow 4 \vdash
((B\rightarrow B) \rightarrow C \rightarrow 5) \rightarrow (4 \rightarrow A) \rightarrow C \rightarrow A$ ($\equiv$
\emph{that}). However, it is not clear what exactly this would mean
for the condition on the \emph{variables}. That no positive (atomic?)
formula can span the empty string?}
\section{Hybrid Type-Logical Grammars}
\label{sec:hybrid}
Hybrid type-logical grammars have been introduced in \cite{kl12gap} as
an extension of lambda grammars which combines insights from the
Lambek calculus into lambda grammars. Depending on authors,
lambda grammars \cite{muskens03lambda} are also called abstract
categorial grammars \cite{groote01acg} or linear
grammars \cite{pollard11linear}.
Formulas of hybrid type-logical grammars are defined as follows, where
$\mathcal{F}_2$ are the formulas of hybrid type-logical grammars and
$\mathcal{F}_1$ the formulas of Lambek grammars. $\mathcal{A}$ denotes
the atomic formulas of the Lambek calculus --- we will call these
formulas \emph{simple} atomic formulas, since their denotations are strings --- $\mathcal{B}$
signifies complex atomic formulas, whose denotations are not simple
strings, but string \emph{tuples}.
\begin{align*}
\mathcal{F}_0 &::= \mathcal{A} \\
\mathcal{F}_1 &::= \mathcal{F}_0 \,\ | \,\ \mathcal{F}_1 / \mathcal{F}_1 \,\
| \,\ \mathcal{F}_1 \backslash \mathcal{F}_1 \\
\mathcal{F}_2 & ::= \mathcal{B} \,\ | \,\ \mathcal{F}_1 \,\ | \,\ \mathcal{F}_2|\mathcal{F}_2
\end{align*}
As is clear from the
recursive definition of formulas above, hybrid type-logical grammars
are a sort of layered or fibred logic. Such logics have been studied
before as extensions of the Lambek calculus by replacing the atomic
formulas in $\mathcal{F}_0$ by feature logic formulas \cite{bj,dm}.
Lambek grammars are obtained by not allowing
connectives or complex atoms in $\mathcal{F}_2$. From hybrid
type-logical grammars, we obtain
lambda grammars by not allowing connectives in
$\mathcal{F}_1$. Inversely, we can see hybrid type-logical grammars as
lambda grammars where simple atomic formulas have been replaced by
Lambek formulas.
Before presenting the rules of hybrid type-logical grammars, we'll
introduce some notational conventions: $A$ and $B$ range over arbitrary formulas; $C$, $D$ and $E$ denote type variables or type constants;
$n$ and $n-1$
denote type constants corresponding to string positions; $\alpha$ and
$\beta$ denote arbitrary types. Types are written as superscripts to
the terms; $x$, $y$ and $z$ denote term variables; $M$ and $N$ denote arbitrary terms.
Table~\ref{tab:htlg} shows the rules of Hybrid Type-Logical
Grammars. The rules are presented in such a way that they compute principal types in
addition to the terms. We obtain the Church-typed version ---
equivalent to the calculus presented in \cite{kl12gap} --- by
replacing all type variables and constants by the type constant
$\sigma$. For the principal types, we use the Curry-typed version,
though for readability, we often write the types of subterms as superscripts as well.
\editout{
$$
\infer[/E]{\lambda z. w_1(\ldots(w_{n+m}\, z)) : A}{\lambda x.
w_1(\ldots(w_n\, x)) : A/B & \lambda y. w_{n+1}(\ldots (w_{n+m}\,
y)) :B }
$$
$$
\infer[\backslash E]{\lambda z. w_1(\ldots(w_{n+m}\, z)) : A}{\lambda x.
w_1(\ldots(w_n\, x)) :B & \lambda y. w_{n+1}(\ldots (w_{n+m}\, y)):B\backslash A}
$$
$$
\infer[/I]{\lambda z. w_1( \ldots (w_{n-1}\, z):A/B}{\infer*{\lambda z. w_1( \ldots w_{n-1}(w_n\, z)):A}{[w_n:B]}}
$$
$$
\infer[\backslash I]{\lambda z. w_2(\ldots (w_n\, z):B\backslash A}{\infer*{\lambda z. w_1( w_2\ldots (w_n\, z)):A}{[w_1:B]}}
$$
$$
\infer[|E]{t\, u:A}{t:A|B & u:B}
$$
$$
\infer[|I]{\lambda x.t:A|B}{\infer*{t:A}{[x:B]}}
$$
The elimination rules are just function composition for type
$\sigma\rightarrow\sigma$. Crucially, the introduction rules have no
representation in the lambda calculus (without term equations).
}
\begin{table}
\begin{center}
\textbf{Lexicon}
\end{center}
\vspace{-.5\baselineskip}
$$
x^{n\rightarrow n-1}: A \vdash M^{\alpha}:A
$$
\begin{center}
\textbf{Axiom/Hypothesis}
\end{center}
\vspace{-.5\baselineskip}
$$
\infer{x^{\alpha}:A \vdash M^{\alpha}:A}{}
$$
\editout{
$$
\infer[/E]{\Gamma,\Delta \vdash \lambda z. w_1(\ldots(w_{n+m}\, z)) :
A}{\Gamma\vdash \lambda x.
w_1(\ldots(w_n\, x)) : A/B & \Delta\vdash\lambda y. w_{n+1}(\ldots (w_{n+m}\,
y)) :B }
$$
$$
\infer[\backslash E]{\Gamma,\Delta\vdash \lambda z. w_1(\ldots(w_{n+m}\, z)) : A}{\Gamma\vdash\lambda x.
w_1(\ldots(w_n\, x)) :B & \Delta\vdash\lambda y. w_{n+1}(\ldots (w_{n+m}\, y)):B\backslash A}
$$
$$
\infer[/I]{\Gamma\vdash\lambda z. w_1( \ldots (w_{n-1}\, z)):A/B}{\Gamma,w_n:B\vdash\lambda z. w_1( \ldots w_{n-1}(w_n\, z)):A}
$$
$$
\infer[\backslash I]{\Gamma\vdash\lambda z. w_2(\ldots (w_n\, z)):B\backslash A}{\Gamma,w_1:B\vdash\lambda z. w_1( w_2\ldots (w_n\, z)):A}
$$
}
\editout{
$$
\infer[/E]{\Gamma,\Delta \vdash (\lambda z^E. M\, (N\, z))^{E\rightarrow C}: A}{\Gamma\vdash M^{D\rightarrow C} : A/B & \Delta\vdash N^{E\rightarrow D} :B }
$$
$$
\infer[\backslash E]{\Gamma,\Delta\vdash (\lambda z^E. M\, (N\, z))^{E\rightarrow
C} :
A}{\Gamma\vdash M^{D\rightarrow C} :B & \Delta\vdash N^{E\rightarrow D}:B\backslash A}
$$
Strictly speaking the rightmost premiss of both rules, which is shown
to be of type $E\rightarrow D$ should be of type $E\rightarrow F$, with the conclusion
of the rule applying the substitution $F := D$ to $\Delta, E$.}
\begin{center}
\textbf{Logical rules -- Lambek}
\end{center}
\vspace{-.5\baselineskip}
$$
\infer[/E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta) \vdash (\lambda z^{\mathbb{s}(E)}. M\, (N\, z))^{\mathbb{s}(E)\rightarrow \mathbb{s}(C)}: A}{\Gamma\vdash M^{F\rightarrow C} : A/B & \Delta\vdash N^{E\rightarrow D} :B }
$$
$$
\infer[\backslash E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta)\vdash (\lambda z^{\mathbb{s}(E)}. M\, (N\, z))^{\mathbb{s}(E)\rightarrow
\mathbb{s}(C)} :
A}{\Gamma\vdash M^{F\rightarrow C} :B & \Delta\vdash N^{E\rightarrow D}:B\backslash A}
$$
$$
\infer[/I]{\mathbb{s}(\Gamma)\vdash ((\lambda x^{\mathbb{s}(D)\rightarrow\mathbb{s}(C)}. M)\, (\lambda z^{\mathbb{s}(F)}.z))^{\mathbb{s}(C)\rightarrow
\mathbb{s}(E)}:A/B}{\Gamma,x^{D\rightarrow C}:B\vdash M^{D\rightarrow E}:A}
$$
$$
\infer[\backslash I]{\mathbb{s}(\Gamma)\vdash ((\lambda x^{\mathbb{s}(C)\rightarrow\mathbb{s}(D)}. M)\, (\lambda z^{\mathbb{s}(F)}.z))^{\mathbb{s}(E)\rightarrow \mathbb{s}(C)}:B\backslash
A}{\Gamma,x^{C\rightarrow D}:B\vdash M^{E\rightarrow D}:A}
$$
\editout{
$$
\infer[/I]{\Gamma\vdash(\lambda z^C. (M[z])^E)^{C\rightarrow
E}:A/B}{\Gamma,x^{D\rightarrow C}:B\vdash (\lambda z^D. (M[(x\, z)])^E)^{D\rightarrow E}:A}
$$
$$
\infer[\backslash I]{\Gamma\vdash (\lambda z^E. M^C)^{E\rightarrow C}:B\backslash
A}{\Gamma,x^{C\rightarrow D}:B\vdash (\lambda z^E. (x\, M^{C})^D)^{E\rightarrow D}:A}
$$
}
\begin{center}
\textbf{Logical rules -- lambda grammars}
\end{center}
\vspace{-.5\baselineskip}
$$
\infer[|E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta)\vdash (M\, N)^{\mathbb{s}(\alpha)} :A}{\Gamma\vdash M^{\beta\rightarrow \alpha}:A|B & \Delta\vdash N^{\gamma}:B}
$$
\editout{
$$
\infer[|E]{\Gamma,\Delta\vdash (M\, N)^{\alpha} :A}{\Gamma\vdash M^{\beta\rightarrow \alpha}:A|B & \Delta\vdash N^{\beta}:B}
$$
As with the other elimination rules, the right premiss of the rule
should have type $\gamma$ and the conclusion of the rule should
compute the most general unifier of $\langle \Gamma; \beta\rangle$ and
$\langle \Delta;\gamma \rangle$ and apply this substitution to the
conclusion of the rule.}
$$
\infer[|I]{\Gamma\vdash (\lambda x^{\beta}.M^{\alpha})^{\beta\rightarrow\alpha}:A|B}{\Gamma,x^{\beta}:B\vdash M^{\alpha}:A}
$$
\caption{Logical rules for hybrid type-logical grammars}
\label{tab:htlg}
\end{table}
The subsystem containing only the rules for $|$ is simply
lambda grammar. The subsystem containing only the rules for $/$ and
$\backslash$ is a notational variant of the Lambek calculus.
For the Lexicon rule, $\langle x^{n\rightarrow n-1}, \alpha\rangle$ is a principal pair for $M$
or, equivalently, with $\lambda x. M$ a $\beta$-normal $\eta$-long linear lambda term
and $(n\rightarrow n-1)\rightarrow\alpha$ its principal type). For the
Axiom/Hypothesis rule, $M$ is the eta-expansion of $x:A$.
For the Lambek calculus elimination rule $/E$ and $\backslash E$,
$\mathbb{s}$ is the most general unifier of $\langle \Gamma;
F\rangle$ and $\langle \Delta; D\rangle$ (this generally just replaces
$F$ by $D$ but takes care of the cases where $C = D$ or $E = F$ as
well). The concatenation operation of the Lambek calculus corresponds
to function composition on terms and to unification of string
positions on types (much like we have seen in Section~\ref{sec:mill1}).
For the Lambek calculus introduction rules $/I$ and $\backslash I$, $\mathbb{s}$ is the most general unifier of $\langle \Gamma;
C\rightarrow D\rangle$ (resp.\ $\langle \Gamma;
D\rightarrow C\rangle$) and $\langle \emptyset; F\rightarrow F\rangle$ (ie.\ we simply identify $C$
and $D$ and replace $x$ by the identity function on string positions
--- the empty string).
In the $|E$ rule, $\mathbb{s}$ is the most general unifier of $\langle
\Gamma;\beta\rangle$ and $\langle \Delta;\gamma\rangle$.
For convenience, we will often tacitly apply the following rule.
$$
\infer[=_{\beta\eta}]{\Gamma \vdash N^{\alpha}:A}{\Gamma \vdash M^{\alpha}:A &
M=_{\beta\eta} N}
$$
Though the above rule is not strictly necessary, we use it to simplify
the lambda terms we compute, performing
on-the-fly $\beta$-normalization (ie.\ we replace $M$ by its
beta-normal, or beta-normal-eta-long, form $N$). Since we have both
subject reduction and subject expansion, $M$ and $N$ are guaranteed to
have the same type $\alpha$.
Apart from the types, the system presented in Table~\ref{tab:htlg} is a
notational variant of hybrid type-logical grammars as presented by
Kubota and Levine \citeyear{kl12gap,kl13emp}. We have replaced strings
as basic types by string \emph{positions}
with Church type $\sigma\rightarrow\sigma$. This is a standard strategy in
abstract categorial grammars, akin to the difference lists in Prolog,
which allows us to do without an explicit concatenation operation:
concatenation is simply treated as function composition, as can be
seen from the term assignments for the $/E$ and $\backslash E$ rules. The
introduction rules $/I$ and $\backslash I$ are presented somewhat differently
than the Kubota and Levine version, who present rules requiring (in
our notation) premisses with term assignments $M \equiv_{\beta\eta} \lambda
z. N[(x\, z)]$ and $M \equiv_{\beta\eta}
(x\, N)$ respectively. The present
formulation has the advantage that it is more robust in the sense that
it does not require us to test that $M$ is $\beta\eta$ equivalent to
the given terms. Though it may appear a bit strange that the $/I$ and
$\backslash I$ rules require the
identity of the type variable $D$ between $x$ and
$M$, it is clear that this follows from the intended interpretation,
which requires the string variable $x$ to occur at the beginning
(resp.\ end) of the string denoted by $M$, and this solution seems preferable to
interleaving normalization and pattern matching in our rules.
The types, at least for the $|$ rules, are exactly those computed using the
principal type algorithm of \citeasnoun{hindley} discussed in Section~\ref{sec:pt}. We will see how the types for the
Lambek connectives and the lexicon rule correspond to principal type
computations in the next section.
\subsection{Justification of the principal types for the new rules}
\label{sec:pthybrid}
For $/E$ and $\backslash E$, their principal types are justified as
follows; $\mathbb{s}_1$ is the most general unifier of $\langle z^G;
G\rangle$ and $\langle \Delta; E\rangle$ --- since $G$ is a type
variable not occurring elsewhere, we can assume without loss of
generality that $\mathbb{s}_1$ just replaces $G$ with $E$ --- and $\mathbb{s}_2$ is the most
general unifier of $\langle
\mathbb{s}_1(\Delta),\mathbb{s}_1(z^G);\mathbb{s}_1(D)\rangle$ and $\langle \Gamma; F\rangle$. The important type unification is of $D$ and
$F$ (the unification of $E$ and $G$ affects only a discharged
axiom).
At the level of the types, the two rules are the same: both correspond
to concatenation.
\editout{
$$
\infer[\rightarrow I]{\lambda z. t\, (u\, z):E\rightarrow C}{
\infer[\rightarrow E]{t\, (u\, z):C}{
\infer[\rightarrow E]{u\, z:D}{
z:E
& u:E\rightarrow D
}
&
t:D\rightarrow C
}
}
$$
}
$$
\infer[\rightarrow I]{\mathbb{s}_2(\Gamma),\mathbb{s}_2(\mathbb{s}_1(\Delta))\vdash \lambda z. M\, (N\, z):\mathbb{s}_2(\mathbb{s}_1(E))\rightarrow \mathbb{s}_2(\mathbb{s}_1(C))}{
\infer[\rightarrow E]{\mathbb{s}_2(\Gamma),\mathbb{s}_2(\mathbb{s}_1(\Delta)), \mathbb{s}_2(\mathbb{s}_1(z^G)) \vdash M\, (N\, z):\mathbb{s}_2(\mathbb{s}_1(C))}{
\infer[\rightarrow E]{
\mathbb{s}_1(\Delta), \mathbb{s}_1(z^G) \vdash N\, z:\mathbb{s}_1(D)}{
\infer{z^G \vdash z:G}{}
& \Delta \vdash N:E\rightarrow D
}
&
\Gamma\vdash M:F\rightarrow C
}
}
$$
Taking $\mathbb{s} = \mathbb{s}_1 \cup \mathbb{s}_2$, which is possible since
$\Gamma$, $\Delta$ and $z$ are disjoint, gives us the following
proof.
$$
\infer[\rightarrow I]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta)\vdash \lambda z. M\, (N\, z):\mathbb{s}(E)\rightarrow \mathbb{s}(C)}{
\infer[\rightarrow E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta), \mathbb{s}(z^G) \vdash M\, (N\, z):\mathbb{s}(C)}{
\infer[\rightarrow E]{
\mathbb{s}(\Delta), \mathbb{s}(z^G) \vdash N\, z:\mathbb{s}(D)}{
\infer{z^G \vdash z:G}{}
& \Delta \vdash N:E\rightarrow D
}
&
\Gamma\vdash M:F\rightarrow C
}
}
$$
Since $\mathbb{s}_1$ only replaced $G$ by $E$ and $G$ no longer
appears in the conclusion of the proof (the corresponding hypothesis
$z$ has been withdrawn) we can treat $\mathbb{s}$ as the most general unifier of
$\langle \Delta;D\rangle$ and $\langle \Gamma;F\rangle$.
\editout{
$$
\infer[\rightarrow I]{\Gamma',\Delta\vdash \lambda z. M\, (N\, z):E\rightarrow C'}{
\infer[\rightarrow E]{\Gamma',\Delta, z^E \vdash M\, (N\, z):C'}{
\infer[\rightarrow E]{
\Delta, z^E \vdash N\, z:D}{
\infer{z^G \vdash z:G}{}
& \Delta \vdash N:E\rightarrow D
}
&
\Gamma\vdash M:F\rightarrow C
}
}
$$
$$
\infer[\rightarrow I]{\Gamma,\Delta\vdash \lambda z. M\, (N\, z):E\rightarrow C}{
\infer[\rightarrow E]{\Gamma,\Delta, z^E \vdash M\, (N\, z):C}{
\infer[\rightarrow E]{
\Delta, z^E \vdash N\, z:D}{
\infer{z^E \vdash z:E}{}
& \Delta \vdash N:E\rightarrow D
}
&
\Gamma\vdash M:D\rightarrow C
}
}
$$
}
We compute the principal type for the $/I$ rule as follows.
$$
\infer[\rightarrow E]{\Gamma \vdash (\lambda x.M) \lambda z.z:C\rightarrow E}{\infer[\rightarrow I]{\Gamma \vdash
\lambda x. M:(D\rightarrow C)\rightarrow D\rightarrow
E}{\Gamma, x^{D\rightarrow C}\vdash M:D\rightarrow E} & \vdash \infer[\rightarrow
I]{\lambda z.z: F\rightarrow F}{\infer{z^F \vdash z:F}{}}}
$$
And symmetrically for $\backslash I$.
$$
\infer[\rightarrow E]{\Gamma \vdash (\lambda x. M)\lambda z.z:E\rightarrow C}{
\infer[\rightarrow I]{\Gamma \vdash \lambda x.M:(C\rightarrow D)\rightarrow E\rightarrow D}{
\Gamma, x^{C\rightarrow D} \vdash M:E\rightarrow D
}
& \infer[\rightarrow I]{\vdash \lambda z.z:F\rightarrow F}{\infer{z^F\vdash z:F}{}}
}
$$
From the point of view of the principal type computation, we identify
the $C$ and $D$ variables, essentially replacing $x$ by the empty string.
\begin{lemma}\label{lemma:pt} The proof rules for Hybrid type-logical
grammars of Table~\ref{tab:htlg} compute principal
types for the lambda terms corresponding to their proofs.
\end{lemma}
\paragraph{Proof} We essentially use the same algorithm as
\citeasnoun{hindley}, which is somewhat simplified by the restriction
to linear lambda terms
which are eta-long.
The principal types for $/ E$, $\backslash E$, $/ I$ and $\backslash I$
rules are justified as shown above.
The lexicon rule is justified by
the Substitution Lemma (Lemma~\ref{lem:subst}): given a principal type $\alpha$ for a lexical entry, we replace a
hypothesis of the form $\alpha \vdash \alpha$ by a hypothesis of the
form $n\rightarrow n-1 \vdash \alpha$, where we know this second sequent has a
linear proof. \hfill \ensuremath{\Box}
\editout{
\begin{lemma}\label{lemma:balanced} If the premisses of a rule in hybrid type-logical
grammars have a balanced type, then so has its conclusion.
\end{lemma}
\paragraph{Proof} This proof is most easily seen using proof
nets. Suppose the premisses of a rule are balanced. Then the premisses
have only a single possible proof net. Now if two atomic formulas are
equal in such a proof net, they must be the conclusions of a single
axiom link. ... TO BE CONTINUED ... NOTE: the key property is that
unified by MGU corresponds to a cut-link (after cut-elimination of the
logical links)
\hfill \ensuremath{\Box}
}
\begin{corollary} Given a principal type derived by
the rules of hybrid type-logical grammar shown above, we can compute
the corresponding lambda term up to
$\beta\eta$ equivalence.
\end{corollary}
\paragraph{Proof}
Since the principal types computed are balanced by Lemma~\ref{lemma:balanced}, by the Coherence theorem (Theorem~\ref{thm:coherence}),
we can compute the corresponding
lambda term up to $\beta\eta$ equivalence. An easy way to do so is to
construct the proof net corresponding to the principal type (which is
unique because of balance) and to compute its lambda term; this
lambda term is the unique beta-normal eta-long term corresponding to
the principal type. \hfill \ensuremath{\Box}
\subsection{Example}
\label{sec:hex}
As an example of how to compute the principal derivation corresponding
to a hybrid derivation, we look at the following hybrid derivation.
{\label{hexproof}
$$
\infer[/I]{\lambda w. (e\, w):s/(np\backslash s)}{\infer[|E]{\lambda
v. (e\, (y\,v)):s}{\infer[|I]{\lambda x\lambda z.(x\, (y\,
z)):s|np}{\infer[\backslash E]{\lambda z.(x\, (y\, z)):s}{[x:np] & [y:np\backslash
s]}} & \lambda P\lambda v. ((P\, e)\, v):s|(s|np) }}
$$
}
\editout{
$$
\infer[???]{e^{1\rightarrow 0} \vdash (\lambda z. e\, z)^{1\rightarrow 0}}{
\infer[\rightarrow E]{e^{1\rightarrow 0}, y^{C\rightarrow 1}\vdash (\lambda z. e\, (y\, z))^{C\rightarrow 0}}{
\infer[\rightarrow I]{y^{C\rightarrow D} \vdash (\lambda x. \lambda z. x\, (y\,
z))^{(D\rightarrow B) \rightarrow C\rightarrow B}}{
\infer[\rightarrow I]{x^{D\rightarrow B}, y^{C\rightarrow D} \vdash (\lambda z. x\, (y\,
z))^{C\rightarrow B}}{
\infer[\rightarrow E]{x^{D\rightarrow B}, y^{C\rightarrow D}, z^C \vdash (x\, (y\, z))^B}{x^{A\rightarrow B} \vdash x^{A\rightarrow B} &
\infer[\rightarrow E]{y^{C\rightarrow D}, z^C \vdash y\, z:D}{y^{C\rightarrow D} \vdash y^{C\rightarrow
D} & z^E \vdash z^E}}}}
&
e^{1\rightarrow 0} \vdash (\lambda P. P\, e)^{((1\rightarrow 0)\rightarrow H\rightarrow G) \rightarrow (H\rightarrow G)}}}
$$
The following works (for the principal type) but is not restrictive
enough (ie.\ overgenerates just like lambda grammars). Require
$C=\textit{max}/\textit{min}$ of the constants in the antecedent?
It seems better to use a test then reduce strategy, that is, test
first that the introduced resource is leftmost/rightmost then perform
the ``standard'' reduction shown below.
}
The corresponding principal derivation looks as follows (for reasons
of vertical space, the lexical entry for $e^{1\rightarrow 0}$ has not been
eta-expanded to $\lambda P \lambda w. (P\, e)\, w$ as it should to
obtain the given principal type instead of $((1\rightarrow 0)\rightarrow G) \rightarrow G$;
though either type will end up being instantiated to the same result type, the eta-expanded principal type $((1\rightarrow 0)\rightarrow H\rightarrow G) \rightarrow
H\rightarrow G$ has the important advantage that it can be obtained without
instantiating type variables to complex types; similarly, $x^{A\rightarrow B}$
and $y^{C\rightarrow D}$ appear in eta-short form).
\medskip
\scalebox{.75}{$$
\infer{e^{1\rightarrow 0} \vdash (\lambda z. e\, z)^{1\rightarrow 0}}{
\infer[\!\rightarrow\!I]{e^{1\rightarrow 0} \vdash (\lambda y. \lambda z. e\, (y\,
z))^{(C\rightarrow 1) \rightarrow C\rightarrow 0}}{
\infer[\!\rightarrow\!E]{e^{1\rightarrow 0}, y^{C\rightarrow 1}\vdash (\lambda z. e\, (y\, z))^{C\rightarrow 0}}{
\infer[\!\rightarrow\!I]{y^{C\rightarrow D} \vdash (\lambda x. \lambda z. x\, (y\,
z))^{(D\rightarrow B) \rightarrow C\rightarrow B}}{
\infer[\!\rightarrow\!I]{x^{D\rightarrow B}, y^{C\rightarrow D} \vdash (\lambda z. x\, (y\,
z))^{C\rightarrow B}}{
\infer[\!\rightarrow\!E]{x^{D\rightarrow B}, y^{C\rightarrow D}, z^C \vdash (x\, (y\, z))^B}{\infer{x^{A\rightarrow B} \vdash x^{A\rightarrow B}}{} &
\infer[\!\rightarrow\!E]{y^{C\rightarrow D}, z^C \vdash (y\, z)^D}{\infer{y^{C\rightarrow D} \vdash y^{C\rightarrow
D}}{} & \infer{z^E \vdash z^E}{}}}}}
&
\!\!\!\!\!\!\!\!\!e^{1\rightarrow 0} \vdash (\lambda P. P\, e)^{((1\rightarrow 0)\rightarrow H\rightarrow G) \rightarrow H\rightarrow
G}}}
& \infer[\!\rightarrow\!I]{\vdash (\lambda v.v)^{J\rightarrow J}}{\infer{v^J \vdash v^J}{}}}
$$}
\medskip
The $\backslash E$ and the $/I$ rules correspond to three rules each in this
principal derivation (the derivation of $\lambda z. x\, (y\, z)$ for
$\backslash E$ and the part of the derivation from $\lambda z. e\, (y\, z)$
to $\lambda z. e\, z$ for $/I$, this last rule satisfies the constraint
for the application of the rule, with $y$ appearing at the last position)
\editout{
Should $C$ be a fresh constant $c$?
$$
\infer[???]{e^{1\rightarrow 0} \vdash (\lambda z. e\, z)^{1\rightarrow 0}}{
\infer[\rightarrow E]{e^{1\rightarrow 0}, y^{c\rightarrow 1}\vdash (\lambda z. e\, (y\, z))^{c\rightarrow 0}}{
\infer[\rightarrow I]{y^{c\rightarrow D} \vdash (\lambda x. \lambda z. x\, (y\,
z))^{(D\rightarrow B) \rightarrow c\rightarrow B}}{
\infer[\rightarrow I]{x^{D\rightarrow B}, y^{c\rightarrow D} \vdash (\lambda z. x\, (y\,
z))^{c\rightarrow B}}{
\infer[\rightarrow E]{x^{D\rightarrow B}, y^{c\rightarrow D}, z^c \vdash (x\, (y\, z))^B}{x^{A\rightarrow B} \vdash x^{A\rightarrow B} &
\infer[\rightarrow E]{y^{c\rightarrow D}, z^c \vdash y\, z:D}{y^{c\rightarrow D} \vdash y^{c\rightarrow
D} & z^E \vdash z^E}}}}
&
e^{1\rightarrow 0} \vdash (\lambda P. P\, e)^{((1\rightarrow 0)\rightarrow H\rightarrow G) \rightarrow (H\rightarrow G)}}}
$$
\editout{
$$
\infer[/I]{\lambda w. (e\, w):1\rightarrow 0\vdash s/(np\backslash s)}{\infer[|E]{\lambda
v. (e\, (y\, v)):c\rightarrow 0\vdash s}{\infer[|I]{\lambda x\lambda z.(x\, (y\,
z)):(1\rightarrow 0)\rightarrow c\rightarrow 0\vdash s|np}{\infer[\backslash E]{\lambda z.(x\, (y\, z)):c\rightarrow 0\vdash s}{[x:1\rightarrow
0\vdash np] & [y:c\rightarrow 1\vdash np\backslash
s]}} & \lambda P\lambda v. ((P\, e)\, v):((1\rightarrow 0)\rightarrow B\rightarrow
A)\rightarrow (B\rightarrow A)\vdash s|(s|np) }}
$$
}
$$
\infer[/I]{\lambda w. (e\, w):1\rightarrow 0\vdash s/(np\backslash s)}{\infer[|E]{\lambda
v. (e\, (y\, v)):c\rightarrow 0\vdash s}{\infer[|I]{\lambda x\lambda z.(x\, (y\,
z)):(1\rightarrow D)\rightarrow c\rightarrow D\vdash s|np}{\infer[\backslash E]{\lambda z.(x\, (y\, z)):c\rightarrow D\vdash s}{[x:C\rightarrow
D\vdash np] & [y:c\rightarrow 1\vdash np\backslash
s]}} & \lambda P\lambda v. ((P\, e)\, v):((1\rightarrow 0)\rightarrow B\rightarrow
A)\rightarrow (B\rightarrow A)\vdash s|(s|np) }}
$$
Can we guarantee that $\Gamma$ never contains variables? Here, the
$/I$ rule guarantees $c$ is a constant. Indeed the fact that we only
combine $\forall$-links with par seems to work in our favour in the Lambek case. However, this
is \emph{not} generally true in the $\lambda$-grammar case, where we
can have par with $\exists$-links. Is this a problem?
}
In principle, the computation of the principal type can fail because of the
constants (even though there might be a proof using
variables). However, this failure would mean the final term fails to
respect the word order of the input string. Principal types using
distinct variables for string positions would seem a useful tool for computing all possible word
orders for a given set of lexical entries, though.
\subsection{Semantics}
\label{sec:sem}
One of the attractive points of categorial grammars is that we have a
very simple and elegant syntax-semantics interface by means of the
Curry-Howard isomorphism between intuitionistic proofs and lambda
terms (or, in our case between linear intuitionistic proofs and linear
lambda terms). By interpreting the logical connectives for the implications ``$/$'', ``$\backslash$'',
``$|$'' and ``$\multimap$'' as the type constructor ``$\rightarrow$'' --- the
\emph{formulas as types} interpretation --- our
derivations in the Lambek calculus, in lambda grammars, in hybrid
type-logical grammars and in first-order linear logic (where we treat
the quantifier as being semantically inert, that is, quantifier rules
are ``invisible'' to the meaning) correspond to $\lambda$-terms ---
the \emph{proofs as terms} interpretation. Using the Curry-Howard
isomorphism, we
can obtain semantics in the tradition of Montague simply by giving
lexical substitutions in the lexicon, using essentially the rules of
Table~\ref{tab:typing} (though we typically use the Church-style
typing) to assign a derivational meaning to a proof.
The semantic version of the proof from the previous section looks as follows.
$$
\infer[/I]{\lambda Q. \forall z. (Q\, z):s/(np\backslash s)}{\infer[|E]{\forall z. (Q\, z):s}{\infer[|I]{\lambda x. (Q\, x):s|np}{\infer[\backslash E]{(Q\, x):s}{[x:np] & [Q:np\backslash
s]}} & \lambda P.\forall z. (P\,z):s|(s|np) }}
$$
Though syntactically, the Lambek elimination rule corresponds to
function composition (concatenation), \emph{semantically} it
corresponds to simple application and the introduction rule to
abstraction. Given the standard Montegovian semantics for ``everyone'' as
$\lambda P.\forall z. (P\, z)$ (the set of properties $P$ such that
all $z$ have this property), the previous proof actually produces an
equivalent term as the semantics for $s/(np\backslash s)$, so the generalized
quantifier can function as a Lambek calculus subject quantifier while
keeping the same semantics.
More detail about the syntax-semantics interface in categorial
grammars can be found in \cite{M95,mr12lcg}.
\section{Equivalence}
\label{sec:equiv}
\editout{
NOTE: this entire section has become superfluous, since the hybrid
proof rules now compute the principal types directly.
NOTE: maybe it is best to separate this into two equivalences: between
hybrid proofs and PT proofs (more precisely, a combination of a MILL
proof, though labeled with hybrid rule names, and a PT proof, where the two proofs have the same structure,
with each
$\multimap E/\multimap I$ matching a $\rightarrow E/\rightarrow I$ and vice versa) and
a PT proof and an MILL1 proof.
\subsection{Hybrid proof to PT proof}
\begin{lemma} Labeled hybrid proofs (that is hybrid proofs with
explicit lambda term labeling) are isomorphic unlabelled hybrid
proofs paired with PT proofs.
\end{lemma}
This lemma is fairly simple, simply exploiting the fact that we can
reconstruct lambda terms from principal types (by Coherence) and vice
versa (by the principal type algorithm).
As a minor technicality, we use $\Gamma \vdash M:\alpha$ instead of
$\vdash M:\alpha$. This is inessential, since we can move back and
forth between $x_1:\alpha_1,\ldots
x_n:\alpha_n \vdash M:\beta$ and $\lambda x_1:\alpha_1 \ldots
x_n:\alpha_n \vdash M^\beta$.
\paragraph{Hybrid to PT}
We simply erase the lambda term of the hybrid proof to produce a
MILL proof with hybrid proof rule names, and erase both the formula
and the lambda term to obtain a proof which corresponds to Hindley's
principal type algorithm (where the Lambek rules are seen as derived
rules as discussed in Section~\ref{sec:pthybrid}), and which retains only
the type information of the hybrid proof. In addition, these two proofs
are linked and have a 1-1 correspondence between formulas in the first
proof and types in the second proof.
We generate this pair of proofs by induction on the depth of the
labeled hybrid proof. In case $d=1$, we have two cases.
\begin{description}
\item{\emph{Lexicon}} $x:A \vdash M:A$, with $M$ a beta-normal
eta-long lambda term of type $\textit{type}(A)$.
For the translation, we compute the MILL type $A'$ corresponding to
$A$ (simply replacing $/$, $\backslash$ and $|$ by $\multimap$) and the two
proofs $A'\vdash A'$ and $PT(\lambda x^{n+1\rightarrow n}.M)$, $n+1 \rightarrow n \vdash
PT(M)$. Since $\lambda x. M$ is linear, $PT(M)$ is balanced.
\item{\emph{Hypothesis}} $x:A \vdash M:A$ with $M$ the eta-expansion of $M:\textit{type}(A)$
For the translation, we simply use $A' \vdash A'$ (again replacing $/$, $\backslash$ and $|$ by $\multimap$) and $\alpha = PT(M)$, $\alpha \vdash \alpha$ (balanced).
\end{description}
In case $d>1$, we proceed by case analysis on the last rule. Induction
hypothesis gives us the required pairs of derivations $\delta_1-\delta_2$ of depth smaller
than $d$ (with $\delta_1$ a MILL proof and $\delta_2$ a PT proof). In
addition, the induction hypothesis guarantees that $\delta_2$ is balanced.
\begin{description}
\item{$[/ E]$} Induction hypothesis gives us proofs
$\delta_1-\delta_3$ of $\Gamma \vdash B\multimap A$ and $\Gamma' \vdash D\rightarrow
C$ respectively and proofs $\delta_2-\delta_4$ of $\Delta \vdash B$
and $\Delta' \vdash E\rightarrow B$ respectively.
$$
\infer[/ E]{\Gamma,\Delta \vdash A}{\infer*[\delta1]{\Gamma \vdash
A/B}{} & \infer*[\delta_2]{\Delta \vdash B}{}}
$$
$$
\infer[\rightarrow I]{\Gamma',\Delta'[B:=D] \vdash E\rightarrow C}{\infer[\rightarrow E,B=D]{\Gamma',\Delta'[B:=D],E\vdash C}{
\infer*[\delta_3]{\Gamma' \vdash D\rightarrow C}{}
& \infer[\rightarrow E,A=E]{\Delta',E\vdash B}{\infer*[\delta_4]{\Delta'\vdash
E\rightarrow B}{} & \infer{A\vdash A}{}}
}}
$$
The net effect of the two substitutions is that the left position of
$E\rightarrow B$ (that is, $B$) is unified with the right position of $D\rightarrow C$
(that is, $D$). Since the principal type computes the type composition
of the lambda terms corresponding to the two premisses, the type of the conclusion of the hybrid
proof rule (as well as its lambda term) match the conclusion of the PT
proof (see
Section~\ref{sec:pthybrid}).
We verify the result is again balanced. Since the premisses of the
rules were balance by induction hypothesis, we know that $\Gamma'$
contains a positive occurrence of $D$ and that $\Delta'$ contains a
negative occurrence of $B$ (which becomes a negative occurrence of $D$
after substitution). For both of these atomic type occurrences, their
respective conjugates (the negative $D$ and the positive $B$) have
disappeared from the conclusion (by the last $\rightarrow E$ rule). Hence, the
conclusion of the rule is again balanced, since it contains no
occurrences of $B$ and both a single positive and a single negative
occurrence of $D$. All other formulas ($E$, $C$) are balanced by
induction hypothesis (being a bit careful in case $C=D$ or $B=E$,
QUESTION: I rather like a proof net argument here, using
$\Gamma'\vdash D\multimap C$, $\Delta'\vdash E\multimap B\ (\text{with}
[B:=D])$ and $E\multimap D, D\multimap C \vdash E\multimap C$ which,
with two cuts becomes $\Gamma',\Delta'\vdash E\multimap C$ where
balance is easy to see, but it's somewhat
inefficient in terms of space. Is it worth adding this?).
\item{$[\backslash E]$} Symmetric. Note that the distinction between the $/ E$ rule
is only in the mapping of the premisses of the hybrid proof to the
left (resp.\ right) premisses of the principal type proof.
\item{$[/ I]$} Induction hypothesis gives us proofs
$\delta_1-\delta_2$ of $\Gamma,B \vdash A$ and $\Gamma',D\rightarrow C\vdash
D\rightarrow E$. We can extend both proofs as follows.
$$
\infer[/I]{\Gamma\vdash A/B}{\infer*[\delta_1]{\Gamma,B\vdash A}{}}
$$
$$
\infer[\rightarrow E]{\Gamma'\vdash C\rightarrow E}{\infer[\rightarrow I]{\Gamma'\vdash (D\rightarrow C)\rightarrow D\rightarrow
E}{\infer*[\delta_2]{\Gamma',D\rightarrow C\vdash D\rightarrow E}{}} & \infer[\rightarrow I]{\vdash F\rightarrow F}{\infer{F\vdash F}{}}}
$$
Where the second proof is the principal type proof for $/I$ of
Section~\ref{sec:pthybrid}. Unlike the Lambek calculus, the current
proof rule does not explicitly exclude
empty antecedent derivations (that is, empty $\Gamma$ with $C=E$ is
not excluded). If needed, we can explicitly require $\Gamma$ to be non-empty.
\item{$[\backslash I]$} Symmetric.
\item{$[| E]$} Induction hypothesis gives us a pair of proofs
$\delta_1-\delta_3$ of $\Gamma \vdash B\multimap A$ and $\Gamma'
\vdash \beta\rightarrow\alpha$ and a pair of proofs $\delta_2-\delta_4$ of
$\Delta\vdash B$ and of $\Delta'\vdash \gamma$. We can combine the
two pairs of proofs as follows.
$$
\infer[| E]{\Gamma,\Delta \vdash A}{\infer*[\delta1]{\Gamma \vdash
B\multimap A}{} & \infer*[\delta_2]{\Delta \vdash B}{}}
$$
$$
\infer[\rightarrow E]{\mathbb{s}(\Gamma'),\mathbb{s} (\Delta') \vdash \mathbb{s} (\alpha)}{\infer*[\delta_3]{\Gamma' \vdash
\beta\rightarrow\alpha}{} & \infer*[\delta_4]{\Delta' \vdash \gamma}{} }
$$
\noindent where $\mathbb{s}$ is the most general unifier of $\langle
\Gamma; \beta\rangle$ and $\langle\Delta; \gamma\rangle$. This is the
same step as case (IVa) of \citeasnoun{hindley}.
\item{$[|I]$} Induction hypothesis gives us a pair of proofs
$\delta_1-\delta_2$ such that $\delta_1$ is a proof of $\Gamma,
B\vdash A$ and such that $\delta_2$ is a proof of $\Gamma', \beta
\vdash \alpha$ where $\beta$ is the type corresponding to $B$.
We can extend both proofs as follows.
$$
\infer[|I]{\Gamma \vdash B\multimap
A}{\infer*[\delta_1]{\Gamma,B\vdash A}{}}
$$
$$
\infer[\rightarrow I]{\Gamma\vdash \beta\rightarrow
\alpha}{\infer*[\delta_2]{\Gamma,\beta\vdash \alpha}{}}
$$
This is case (II) of \citeasnoun{hindley}.
\end{description}
\paragraph{PT to Hybrid}
IDEA: use Coherence (since all types are balanced) covert PT to lambda
term (replacing all atomic types by $\sigma$).
Lexicon: $n+1 \rightarrow n \vdash \alpha$, MILL1 proof $A\vdash A$, we know by Coherence there is a
unique beta-normal eta-long lambda term $\lambda x. M$ which corresponds to the balanced
principal type $(n+1 \rightarrow n) \rightarrow \alpha$. $x:A \vdash M:A$
$\multimap E/\rightarrow E$:}
For the main result, we only need to show that a hybrid principal type
proof corresponds to a MILL1 proof, since we can reconstruct the
lambda term from the principal type.
The basic idea which makes the correspondence work is that there is a
1-1 mapping between the atomic terms of a predicate in MILL1 and the
principal type which is assigned to the corresponding term in a hybrid derivation.
So from the term assigned to a hybrid derivation, we compute the
principal type using the principal type algorithm (PTA) and this gives us
the first-order variables and from the first-order variables of a
MILL1 derivation we obtain the principal type and a hybrid lambda term
thanks to the coherence theorem, as shown schematically below.
\begin{tikzpicture}
\node (lt) {Hybrid lambda term};
\node (pt) [right=3em of lt] {Principal type};
\node (fo) [right=3em of pt] {First-order variables};
\draw[<->] (fo) -- (pt);
\path[->] ([yshift=+2pt] lt.east) edge node[above] {PTA} ([yshift=+2pt] pt.west);
\path[<-] ([yshift=-2pt] lt.east) edge node[below] {Coherence} ([yshift=-2pt] pt.west);
\end{tikzpicture}
\subsection{String positions, types and formulas}
\label{sec:string}
We need an auxiliary function $f$ (for \emph{flatten}) which reduces a complex type to a list
of atomic types. Following \citeasnoun{kanazawa11pg}, we compute this list by first taking the yield
of the type tree and then reversing this list, which is convenient for induction
since it has $f(\beta\rightarrow \alpha) = f(\alpha)\concat f(\beta)$ ( ``$\concat$''
denotes list concatenation, $[ A ]$ the singleton list containing
element $A$ and $[ A_1,\ldots,A_n ]$ the $n$-element list with $i$th
element $A_i$).
\begin{definition}\label{def:flatten} Let $\alpha$ be a type, the list $f(\alpha)$ is
defined as follows.
\begin{align*}
f(A) &= [ A ] \quad \text{when $A$ atomic} \\
f(\beta\rightarrow \alpha) &= f(\alpha)\concat f(\beta)
\end{align*}
\end{definition}
For example, we have the following.
$$
\begin{array}{l}
f((B\rightarrow 2) \rightarrow (1\rightarrow A) \rightarrow B\rightarrow A) \\
= f((1\rightarrow A) \rightarrow B\rightarrow A) \concat f(B\rightarrow 2) \\
= f(B\rightarrow A) \concat f(1\rightarrow A) \concat f(B \rightarrow 2) \\
= [ A, B, A, 1, 2, B ]
\end{array}
$$
\editout{
\begin{definition} The length of a type $\alpha$ is defined as
follows.
\begin{align*}
\textit{length}(A) & = 1 \\
\textit{length}(\beta \rightarrow \alpha) &= \textit{length}(\alpha) +
\textit{length}(\beta)
\end{align*}
\end{definition}
Since the length of type $\alpha$ simply counts the number of leaves
and the function $f(\alpha)$ collects the leaves of $\alpha$ (in
right-to-left order) in a list, the length of the type $\alpha$ corresponds to
the length of the list $f(\alpha)$. Given a type $\alpha$ and a list
of types $L$
of the same length, we can instantiate $\alpha$ with $L$ as follows.
\begin{definition} Given a list of atomic type variables/constants and a type of the same length.
\begin{align*}
f^{-1}([A],\alpha) &= A \\
f^{-1}([A_1,\ldots,A_n],\beta\rightarrow\alpha) &=
f^{-1}([A_{i+1},\ldots,A_n],\beta) \rightarrow f^{-1}([A_1,\ldots,A_i],\alpha)
\\ &
\qquad \qquad \text{where $i$ is length($\alpha$)}
\end{align*}
\end{definition}
From the definition, it is immediate that $f^{-1}(\alpha)$ is always
defined if the leaves of $\alpha$ are distinct type variables disjoint
from those in $L$.
As is suggested by the notation $f^{-1}$ functions as a sort of
inverse to $f$, allowing us to reconstruct a type from the list of
atomic formulas produced by $f$ plus the structure of the original
type. As an example, given the type $(C\rightarrow D) \rightarrow (E\rightarrow F) \rightarrow G\rightarrow H$ and
the list of atomic types $[ A, B, A, 1, 2, B ]$.
$$
\begin{array}{l}
f^{-1}([ A, B, A, 1, 2, B ], (C\rightarrow D) \rightarrow (E\rightarrow F) \rightarrow G\rightarrow H) \\
= f^{-1}(C\rightarrow D,[2, B]) \rightarrow f^{-1}((E\rightarrow F)\rightarrow G\rightarrow H, [A, B, A, 1])
\end{array}
$$
TODO: complete
}
\editout{
Given such a list and a term respecting the structure of the
initial term where all leaves are distinct variables (in our case
$(C\rightarrow D) \rightarrow (E\rightarrow F) \rightarrow G\rightarrow H$), there is a
unique instantiation of this term with this list of variables which
reconstructs the original term, simply by instantiating the variables
in this term from left-to-right with the atoms in the list from right-to-left (for our example this produces
the instantiation $C=B$, $D=2$, $E=1$, $F=A$, $G=B$, $H=A$).
}
\editout{
However, when converting these lists to the arguments of atomic
formulas at the end of the recursion, I prefer using the following
function $a$ for
flattening, which has the advantage of producing the list in a
convient left-to-right order (respecting this left-to-right order
might require some permutation of the arguments). This strategy can be exploited for
non-well-nested grammar as well, by passing through a tuple-forming stage.
$$
\begin{array}{rl}
a(B\rightarrow A) &= [ A,B ] \\
a((C\rightarrow B) \rightarrow (D\rightarrow A)) &= [ A,B,C,D ] \\
a((A_2 \rightarrow A_1) \rightarrow \ldots (A_{n-1} \rightarrow A_{n-2}) \rightarrow A_n \rightarrow A_0)
&= [ A_0,A_1,A_2,\ldots,A_{n-2},A_{n-1},A_n ] \\
\end{array}
$$
}
\editout{
Formulas are translated as follows.
$$
\begin{array}{ll}
\| p \|^{[C_1,\ldots,C_n]} & = p(C_1,\ldots,C_n) \\
\| A|B \|^{f(\beta\rightarrow\alpha)} &= \| B \|^{f(\beta)} \multimap \| A
\|^{f(\alpha)} \\
\| A/B \|^{[ C,D ]} &= \forall x. \| B \|^{[ D,x ]} \multimap \| A
\|^{[ C,x]} \\
\| B\backslash A \|^{[ C,D ]} &= \forall x. \| B \|^{[ x,C ]} \multimap \| A
\|^{[ x,D]} \\
\end{array}
$$
As a final translation step, we can either obtain the explicitly
quantified translation of a formula, by universally quantifying over
all free variables in the formula or the implicitly quantified version
by removing all negative occurrences of $\forall$. For the equivalence
proof, we will use the implicitly quantified version.
An alternative way to define the explicitly quantified rules would be
to change the $A|B$ rule to quantify over the variables in the intersection of
$f(\alpha)$ and $f(\beta)$ (translating both lists into sets)
and to change the atom rule to quantify over any remaining free
variables.
More precisely, we have the following two translations from polarized
formulas to implicitly quantified and to explicitly quantified MILL1 formulas.
Implicitly quantified:
}
\begin{definition}\label{def:transl} Let $A$ be a formula in Hybrid Type-Logical
Grammar, $\alpha$ its principal type and $L = f(\alpha)$ the
flattened list of atomic types obtained from $\alpha$ according to Definition~\ref{def:flatten}. The translation of $A$ into
first-order linear logic is defined as follows.
$$
\begin{array}{ll}
\| p \|^{[C_1,\ldots,C_n]} & = p(C_1,\ldots,C_n) \\
\| (A|B) \|^{f(\beta\rightarrow\alpha)} &= \| B \|^{f(\beta)} \multimap \| A
\|^{f(\alpha)} \\
\| (A/B) \|^{[ C,D ]} &= \forall x. \| B \|^{[ D,x ]} \multimap \| A
\|^{[ C,x]} \\
\| (B\backslash A) \|^{[ C,D ]} &= \forall x. \| B \|^{[ x,C ]} \multimap \| A
\|^{[ x,D]} \\
\end{array}
$$
We can obtain a closed formula by universally quantifying over all
variables in the list of arguments replacing all of them with
quantified variables using the universal closure operation (Definition~\ref{def:cl}).
$$
\| A \|_c^{L} = \textit{Cl}(\| A \|^L)
$$
\end{definition}
\begin{proposition}\label{prop:fvt} Let $A$ be a formula in first-order linear logic
and $H$ a formula in hybrid type-logical grammar and $A \equiv \| H \|^{f(\alpha)}$. The free
meta-variables of $A$ are exactly the type variables of $\alpha$
(and of $f(\alpha)$).
\end{proposition}
\paragraph{Proof} Immediate by induction on $H$ using the
translation. All new variables introduced during the translation are
bound. \hfill \ensuremath{\Box}
\begin{lemma}\label{lem:mgu} Let $A_1$ and $A_2$ be first-order linear logic formulas
obtained by the translation function from Hybrid Type-Logical
Grammar formulas $H_1$ and $H_2$ with
$\gamma_1$ and $\gamma_2$ as their respective principal types. In
other words, $A_1 \equiv \| H_1 \|^{f(\gamma_1)}$ and $A_2 \equiv \| H_2 \|^{f(\gamma_2)}$.
$A_1$ unifies with $A_2$ with MGU $\mathbb{s}$ if and only if $H_1 \equiv
H_2$ and $\gamma_1$ unifies with $\gamma_2$ with this same MGU $\mathbb{s}$.
\end{lemma}
\paragraph{Proof} Suppose $A_1$ and
$A_2$ unify with MGU $\mathbb{s}$. We must show that $H_1 \equiv H_2$ and
that $\mathbb{s}$ is an MGU for $\gamma_1$ and $\gamma_2$. Showing $H_1 \equiv H_2$
is an easy induction (exploiting the fact that $A|B$ does not have a
quantifier prefix and therefore cannot unify with a Lambek connective
and that $A/B$ and $B\backslash A$ cannot unify with each other because of the
condition preventing accidental capture of variables). Given that
$H_1$ and $H_2$ are identical, we know that $A_1$ and $A_2$ differ
only in the free variables (the bound variables are equivalent up to
renaming) and that the free variables for $A_1$ and $A_2$ are exactly
the type variables of $\gamma_1$ and $\gamma_2$ (by
Proposition~\ref{prop:fvt}). Therefore any substitution that makes
$A_1$ and $A_2$ equal (up to renaming of bound variables) makes
$\gamma_1$ and $\gamma_2$ equal.
For the other direction, suppose that $H_1 \equiv H_2$ and that
$\mathbb{s}$ is the MGU of $\gamma_1$ and $\gamma_2$. Since $\mathbb{s}$ is a MGU
$\mathbb{s}(\gamma_1) \equiv \mathbb{s}(\gamma_2)$ and therefore given that the translation
function uses identical hybrid formulas and identical principal types
we have that
$A_1 \equiv A_2$.
\hfill \ensuremath{\Box}
\editout{
Explicitly quantified:
$$
\begin{array}{ll}
\| p \|^{+,[C_1,\ldots,C_n]} & = \forall x_1,\ldots,x_m. p(C_1,\ldots,C_n) \\
\| p \|^{-,[C_1,\ldots,C_n]} & = \forall x_1,\ldots,x_m. p(C_1,\ldots,C_n) \\
\| (A|B) \|^{+,f(\beta\rightarrow\alpha)} &= \exists x_1,\ldots,x_m. \| B \|^{-,f(\beta)} \multimap \| A
\|^{+,f(\alpha)} \\
\| (A|B) \|^{-,f(\beta\rightarrow\alpha)} &= \forall x_1,\ldots,x_m. \| B \|^{+,f(\beta)} \multimap \| A
\|^{-,f(\alpha)} \\
\| (A/B) \|^{+,[ C,D ]} &= \forall x. \| B \|^{-,[ D,x ]} \multimap \| A
\|^{+,[ C,x]} \\
\| (A/B) \|^{-,[ C,D ]} &= \forall x. \| B \|^{+,[ D,x ]} \multimap \| A
\|^{-,[ C,x]} \\
\| (B\backslash A) \|^{+,[ C,D ]} &= \forall x. \| B \|^{-,[ x,C ]} \multimap \| A
\|^{+,[ x,D]} \\
\| (B\backslash A) \|^{-,[ D,E ]} &= \forall x. \| B \|^{+,[ x,D ]} \multimap \| A
\|^{-,[ x,E]} \\
\end{array}
$$
Let $Z$ be the set of free variables used at the start of the
translation (remember that non-lexical axioms start with free
variables).
For the $(A|B)$ case, $x_1,\ldots,x_m$ is the
set of free variables in $f(\beta) \cap f(\alpha)$ \emph{minus} the
free variables in $Z$; for the atomic
formulas, $x_1,\ldots,x_m$ is the set of free variables in
$C_1,\ldots,C_n$ minus the free variables in $Z$).
}
It is insightful to compare the translation of $(np\backslash s)/np$ (with
principal type $2\rightarrow 1$) to that of $(s|np)|np$ with principal type $(B\rightarrow
2) \rightarrow (1\rightarrow A) \rightarrow B\rightarrow A$. Though the two end results are formulas
which are equivalent to each other (after universal closure of the meta-variables), there is a difference in the string position list for
the non-atomic subformulas: the Lambek formula only ever has a pair of
string positions, whereas the linear formula starts with a full list
of string positions which decreases at each step. In other words, for the Lambek formula, we compute the string positions step-by-step
whereas the lambda grammar version of the same formula precomputes all
string positions then divides them among the subformulas.
$$
\begin{array}{l}
\| (np\backslash s)/np \|^{[1,2]}
\\
= \forall y. \| np \|^{[2,y]} \multimap \| np \backslash s \|^{[1,y]}
\\
= \forall y. np(2,y) \multimap \| np \backslash s \|^{[1,y]}
\\ = \forall y. np(2,y) \multimap \forall x. \| np \|^{[x,1]} \multimap \| s \|^{[x,y]}
\\ = \forall y. np(2,y) \multimap \forall x. [ np(x,1) \multimap
s(x,y) ]
\end{array}
$$
$$
\begin{array}{l}
\| (s|np)|np \|^{[ A, B, A, 1, 2, B ]} \\
= \| np \|^{[2,B]} \multimap \| s|np \|^{[ A, B, A, 1]} \\
= np(2,B) \multimap \| s|np \|^{[ A, B, A, 1]} \\
= np(2,B) \multimap \| np \|^{[A,1]} \multimap \| s \|^{[ A, B]} \\
= np(2,B) \multimap np(A,1) \multimap s(A,B) \\
\\
\| (s|np)|np \|_c^{[ A, B, A, 1, 2, B ]} \\
= \forall x. \forall y. \{ np(2,B) \multimap np(A,1) \multimap s(A,B)\}[A:=x,B:=y] \\
= \forall x. \forall y. [np(2,y) \multimap np(x,1) \multimap s(x,y)] \\
\end{array}
$$
\editout{
$$
\begin{array}{l}
\| np \rightarrow np\rightarrow s \|^{[ A, B, A, 1, 2, B ]} \\
= \| np \|^{[2,B]} \rightarrow \| s|np \|^{[ A, B, A, 1]} \\
= (B\rightarrow 2) \rightarrow \| np\rightarrow s \|^{[ A, B, A, 1]} \\
= (B\rightarrow 2) \rightarrow \| np \|^{[A,1]} \rightarrow \| s \|^{[ A, B]} \\
= (B\rightarrow 2) \rightarrow (1\rightarrow A) \rightarrow (B\rightarrow A) \\
\end{array}
$$
}
\editout{
There is an important difference therefore in the translation of the
axiom sequent for $(np\backslash s)/np$ (with input positions 1,2) and
for $(s|np)|np$ (with sequence of types $[A,B,A,1,2,B]$): the former
produces sequent~\ref{seqone} whereas the second produces
sequent~\ref{seqtwo} below. The former sequent is of the right form
for a $\forall I$ rule, whereas the second, due to the free
occurrences of $A$ and $B$ on the left hand side of the turnstile, is not.
\begin{align}
\label{seqone} \forall y. np(2,y) \multimap \forall x. [ np(x,1) \multimap
s(x,y) ] & \vdash np(2,B) \multimap np(A,1) \multimap s(A,B) \\
\label{seqtwo} np(2,B) \multimap np(A,1) \multimap s(A,B) & \vdash np(2,B) \multimap
np(A,1) \multimap s(A,B)
\end{align}}
\editout{
The following lemma gives us a generalization of the focus shift rule
for natural deduction. The proof
essentially gives us an version of eta-expansion specialized for the translated
formulas of hybrid type-logical grammars and shows that the positive
and negative translation functions are well-behaved with respect to
the each other and the derivability relation.
\begin{lemma} We can add the following derived rule for translated
sequents to first-order linear logic.
$$
\infer[pn]{\Gamma \vdash_p \| C \|^{+,L}}{\Gamma
\vdash_n \| C \|^{-,L}}
$$
Where $\Gamma$ is any context (of first-order linear logic), $C$ any formula (of hybrid type-logical
grammar) and $L$ any appropriate list of
string positions for the translation function.
\end{lemma}
\paragraph{Proof} By induction on the structure of $C$.
If $C$ is atomic, $\| C \|^+$ and $\| C \|^-$ are identical and we can
simply apply the focus shift rule.
If $C$ is of the form $B | A$ then we know by induction
hypothesis that the rule is valid for $A$ and $B$, which we can use as follows.
$$
\infer[=_{\textit{def}}]{\Gamma \vdash_p \| B | A \|^{+,f(\alpha\rightarrow\beta)}}{\infer[\multimap I]{\Gamma\vdash_p \| A \|^{-,f(\alpha)} \multimap \| B \|^{+,f(\beta)}}{\infer[IH]{\Gamma,\| A \|^{-,f(\alpha)}
\vdash_p \| B \|^{+,f(\beta)}}{\infer[\multimap E]{\Gamma,\| A \|^{-,f(\alpha)} \vdash_n \|
B \|^{-,f(\beta)}}{\infer[=_{\textit{def}}]{\Gamma \vdash_n \| A
\|^{+,f(\alpha)} \multimap \| B \|^{-,f(\beta)}}{\Gamma \vdash_n
\| B|A \|^{-,f(\alpha\rightarrow\beta)}} & \infer[IH]{\| A \|^{-,f(\alpha)}
\vdash_p \| A \|^{+,f(\alpha)}}{\| A \|^{-,f(\alpha)} \vdash_n \| A \|^{-,f(\alpha)}}}}}}
$$
If $C$ is of the form $B/A$ then we proceed in a similar way, the only
difference being that we use the fact that the meta-variable $E$ does
not occur in $\Gamma$ to allow us to use the $\forall I$ rule.
$$
\infer[=_{\textit{def}}]{\Gamma \vdash_p \| B/A
\|^{+,[C,D]}}{\infer[\forall I]{\Gamma\vdash_p \forall x. \| A \|^{-,[D,x]} \multimap \| B \|^{+,[C,x]}}{\infer[\multimap I]{\Gamma\vdash_p \| A \|^{-,[D,E]} \multimap \| B \|^{+,[C,E]}}{\infer[IH]{\Gamma,\| A \|^{-,[D,E]}
\vdash_p \| B \|^{+,[C,E]}}{\infer[\multimap E]{\Gamma,\| A \|^{-,[D,E]} \vdash_n \|
B \|^{-,[C,E]}}{\infer[=_{\textit{def}}]{\Gamma \vdash_n \| A
\|^{+,[D,E]} \multimap \| B \|^{-,[C,E]}}{\Gamma \vdash_n
\| B/A \|^{-,[C,D]}} & \infer[IH]{\| A \|^{-,[D,E]}
\vdash_p \| A \|^{+,[D,E]}}{\| A \|^{-,[D,E]} \vdash_n \| A \|^{-,[D,E]}}}}}}}
$$
The case for $C \equiv A\backslash B$ is symmetric.
\hfill \ensuremath{\Box}
}
Remember that sequents in hybrid type-logical grammar are of the form $x_1^{\alpha_1}:A_1, \ldots,
x_n^{\alpha_n}:A_n \vdash M^{\beta}:B$ with
$M$ a linear lambda term containing exactly the free variables $x_1,
\ldots, x_n$ and that the principal type of $\lambda x_1,
\ldots x_n . M$ is balanced and of the form $\alpha_1 \rightarrow \ldots
\rightarrow a_n
\rightarrow \beta$. For the translation, we separate lexical axioms
from other axioms: lexical axioms correspond to closed formulas,
whereas the other axioms typically have free variables. With this in mind, we translate sequents as $\| A_1
\|^{f(\alpha_1)}, \ldots, \| A_n
\|^{f(\alpha_n)} \vdash \| B
\|^{f(\beta)}$, where the translation $\| . \|_c$ is used for
hypotheses which start at a lexicon rule and $\| . \|$ for hypothesis
which start at the axiom rule. For the right-hand side $B$, we use the
universal closure of all free variables in $f(\beta)$ \emph{minus} the
free variables on the left hand side of the sequent (the only free
variables are those used in the translation of hypothesis rules), this
is the universal closure of $B$ modulo $\Gamma$ of Definition~\ref{def:cl}.
In order not to overburden our notation,
when the types are understood from the context, we
will often abbreviate this translation as $\| \Gamma\|
\vdash \| B \|$ (or even as $\Gamma\vdash \| B \|$,
leaving the translation of $\Gamma$ implicit). As a special case of this translation, the sequent $w_1^{1\rightarrow 0}:A_1, \ldots
w_n^{n\rightarrow n-1}:A_n \vdash M^{n\rightarrow 0}:B$, which is the endsequent
corresponding to a sentence in a hybrid type-logical grammars, is translated as $\|
A_1 \|_c^{[0,1]},\ldots, \| A_n \|_c^{[n-1,n]}
\vdash \| B \|_c^{[0,n]}$.
\editout{
$$
\infer*[[\forall I]^*]{}{\infer*[[\multimap
I]^*]{}{\infer*[[\multimap I/\forall
I]^*]{}{\infer[\pm]{}{\infer[[\forall E/\multimap
E]]{}{\infer*{}{\infer[\forall E/\multimap E]{}{}} & }}}}}
$$
}
\paragraph{Example: gapping} To give an example, the gapping lexical entry for ``and'' of
\cite{kl12gap} looks as follows in our notation.
$$
((s|tv)|(s|tv))|(s|tv): \lambda \textit{STV2} \lambda \textit{STV1} \lambda \textit{TV} \lambda z.
(\textit{STV1}\ \textit{TV}) (\textit{and}\ (\textit{STV2}\ \lambda x.x))
$$
\noindent where $tv$ is short for $(np\backslash s)/np$. The principal type
for this lambda term would be the following (the corresponding
formulas have been annotated above for ease of comparison).
$$
((\overset{tv}{\overbrace{E\rightarrow E}})\rightarrow \overset{s}{\overbrace{D\rightarrow 4}}) \rightarrow ((\overset{tv}{\overbrace{B\rightarrow A}}) \rightarrow \overset{s}{\overbrace{3\rightarrow C}}) \rightarrow (\overset{tv}{\overbrace{B\rightarrow A}}) \rightarrow \overset{s}{\overbrace{D\rightarrow C}}
$$
If $tv$ were an atomic formula, the first-order linear logic formula
would look as shown below on the first line, the complete formula (for
the positive translation) is
shown just below it.
\begin{gather*}
(tv(E,E) \multimap s(4,D)) \multimap (tv(A,B) \multimap s(C,3))
\multimap tv(A,B) \multimap s(C,D) \equiv \\
(\forall v. [ np(v,E) \multimap \forall w. [ np(E,w) \multimap s(v,w)] ] \multimap s(4,D))
\multimap \\ (\forall x' [ np(x,A) \multimap \forall y' [ np(B,y')
\multimap s(x',y') ] ]\multimap s(C,3)) \multimap \\
\qquad\qquad\qquad \forall x. [ np(x,A) \multimap \forall y. [
np(B,y) \multimap s(x,y) ] ]
\multimap s(C,D)
\end{gather*}
Though the formula above looks intimidating (even before universal closure), it is easy to verify that it is \emph{equivalent} (up
to variable names) to the first-order linear logic formula which
corresponds to the analysis of gapping for the Displacement calculus
from Section~3.2.6 of \cite{mvf11displacement},
using the translation given in \cite{moot13lambek}.
\subsection{Proof-theoretic properties of the translation into MILL1}
Before proving the main theorem, stating that for every hybrid proof
there is a first-order linear logic proof of its translation, we will
spend some time on the structure of normal/focused natural deduction
proofs and the consequences of the translation function.
Given that in hybrid type-logical grammars, the
lambda-grammar connective ``$|$'' always outscopes the Lambek
connectives ``$/$'' and ``$\backslash$'', proofs using
the translated formulas into focused first-order linear logic look
schematically as shown in Figure~\ref{fig:transl}.
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (lsp) at (7.6,4.2) {Subproofs $\Delta_i \vdash_p \| \mathcal{F}_1 \|$};
\node (hsp) at (6.5,5.25) {Subproofs $\Gamma_j \vdash_p \| \mathcal{F}_2 \|$};
\fill[black!10] (0,1.85) rectangle (6.1,2.75);
\draw[fill=black!10,black!10] (0,3.25) -- (6.1,3.25) -- (5.2,4.15) -- (0,4.15) -- (0,3.25);
\draw[dotted] (6.1,3.35) -- (6.4,3.65);
\draw[dotted] (5.7,3.75) -- (6.0,4.05);
\draw[dotted] (5.3,4.15) -- (5.6,4.45);
\draw[dotted] (4.95,4.5) -- (5.25,4.8);
\draw[dotted] (4.65,4.8) -- (4.95,5.1);
\draw[dotted] (4.4,5.05) -- (4.7,5.35);
\node (bottom) at (6,-0.2) {7};
\node (ii) at (6,0.8) {6};
\node (lli) at (6,1.8) {5};
\node (sf) at (5.75,3) {$\pm$\ \ 4};
\node (sft) at (6,3.2) {};
\node (sfb) at (6,2.8) {};
\node (le) at (5,4.2) {3};
\node (lbe) at (4,5.2) {2};
\node (fe) at (4,6.2) {1};
\path (bottom) edge node[left] {$[\forall I]^*$} (ii);
\path (ii) edge node[left] {$[\multimap I]^*$} (lli);
\path (lli) edge node [left] {$[\multimap I/\forall I]^*$} (sfb);
\path (sft) edge node [left] {$\ \ [\forall E/\multimap E]^*$} (le);
\path (le) edge node [left] {$\ \ [\multimap E]^*$} (lbe);
\path (lbe) edge node [left] {$[\forall E]^*$} (fe);
\node[text width=7.5em] (lambdai) at (1,0.8) {\hfill $\lambda$ grammar I $\left\{ \rule{0pt}{2.6em}\right.$};
\node[text width=7.5em] (lambeki) at (1,2.3) {\hfill Lambek I $\left\{ \rule{0pt}{1.4em}\right.$};
\node[text width=7.5em] (lambeke) at (1,3.7) {\hfill Lambek E $\left\{ \rule{0pt}{1.4em}\right.$};
\node[text width=7.5em] (lambdai) at (1,5.1) {\hfill $\lambda$ grammar E $\left\{
\rule{0pt}{2.6em}\right.$};
\node[text width=7.5em] (shfocus) at (1,3){\hfill focus shift $\{\,$};
\node[text width=7.5em] (lexicon) at (1,6.2){\hfill lexicon/axiom $\{\,$};
\end{tikzpicture}
\end{center}
\caption{Schematic form of the main track of a translated hybrid
sequent.}
\label{fig:transl}
\end{figure}
The figure shows the main track of a proof, which starts either with a
hypothesis/axioms, then has an elimination part, followed by a focus
shift followed by an introduction part ending in the conclusion of the
proof --- this is just the definition of a main track
(Definition~\ref{def:track}). The definition of formulas guarantees that the elimination part
starts with any number of $[| E]$ rules (possibly zero, like all other
parts, as indicated by the $*$ superscript in the figure) followed by
any combination of $[/ E]$ and $[\backslash E]$ rules. The order is inverse
in the introduction part of the track, with $[/ I]$ and $[\backslash I]$
preceding $[| I]$. For the translation of these rules into first-order
linear logic, the quantifiers corresponding to the
rules for the lambda grammar connective ``$|$'' are obtained by
universal closure, so if they are present, it must be as a prefix at
the beginning of the proof or as a postfix at the end of the proof
--- the subpaths labeled (1)-(2) and (6)-(7) in
Figure~\ref{fig:transl} --- and the Lambek connectives correspond to
a combination of a $\forall$ and a $\multimap$ rule upon translation.
\begin{proposition}\label{prop:forall} \emph{a.}\ The main track of a
focused proof of a translated hybrid sequent looks as shown in Figure~\ref{fig:transl}.
\emph{b.}\ The subproofs $\Gamma_j \vdash_p \| \mathcal{F}_2
\|$ contain the sequence of proof steps in (1)-(6), that is they do
not end with any $\forall I$ rules corresponding to a hybrid
connective.
\emph{c.}\ The subproofs in $\Delta_i \vdash_p \| \mathcal{F}_1 \|$ contain the
sequence of proof steps in (1)-(5), that is they do not end with any
lambda grammar introduction rules.
\end{proposition}
\paragraph{Proof} These are immediate consequences of the translation function and the
structure of normal proofs.
\emph{a.}\ follows from the way the translation function is defined
and the standard structure of a main track.
\emph{b.}\ since normal proofs satisfy the subformula property and since hybrid connectives
are translated into prenex formulas, we do
not produce subformulas of the form $(\forall x_1,\ldots, x_n. [ A
\multimap B ]) \multimap C$ (for $n\geq 1$).
\emph{c.}\ would contradict the
definition of hybrid formulas, since it would have a Lambek connective
outscope a lambda grammar connective. \hfill \ensuremath{\Box}
An immediate corollary of Proposition~\ref{prop:forall} is that $\forall I$ rules corresponding to lambda
grammar connectives occur only at the end of the main track of a
proof, just like $\forall E$ rules corresponding to lambda grammar
connective occur only at the start of any track in which they occur.
\begin{definition} We say a first-order linear logic proof obtained by
translating a hybrid proof is in \emph{quantifier-reduced form},
when all $\forall E$ and $\forall I$ rules obtained by universal closure of
lambda-grammar connectives have been removed from the proof.
More precisely, the translation is kept as before with the following
two exceptions:
\begin{itemize}
\item the Lexicon rule is translated as $\| A \|_c^{f(\alpha)}
\vdash \| A \|^{f(\alpha)}$ (with the closure operation applied only to
the translation of the antecedent)
\item the endsequent is translated as
$\Gamma \vdash \| C \|^{f(\beta)}$ (without the usual closure modulo
$\Gamma$).
\end{itemize}
\end{definition}
\begin{proposition} A sequent $\Gamma \vdash A$ produced by the
translation function is derivable if and only if its
quantifier-reduced form is.
\end{proposition}
\paragraph{Proof} Immediate by Proposition~\ref{prop:cl}. \hfill \ensuremath{\Box}
Quantifier-reduced form is a way of ``compiling'' away the predictable
prefixes of $\forall E$ rules (for each of the lexical leaves of the
proof) and the equally predictable postfix of
$\forall I$ rules introduced by the universal closure operation.
This
simplifies the structure of the proof, as is clear from
Proposition~\ref{prop:qr} below and from
Figure~\ref{fig:transl} --- we keep only the subpath (2)-(6). It also
simplifies the correctness proof of the translation in the following
sections, since we avoid having to start each inductive step by a
number of $\forall E$ rules and end it with a number of $\forall I$ rules.
The quantifier-reduced
form of a proof is sensitive to the way we have obtained the formula:
the Lexicon rule for the Lambek formula $np\backslash s$ has quantifier-reduced form $\forall
x. np(x,1)\multimap s(x,2)\vdash_n \forall
y. np(y,1)\multimap s(y,2)$ whereas Lexicon rule for the formula $s|np$ with principal
type $(1\rightarrow A) \rightarrow (2\rightarrow A)$, which would normally be assigned the
same axiom, has quantifier-reduced form $\forall
x. np(x,1)\multimap s(x,2)\vdash_n np(A,1)\multimap s(A,2)$ (which we
can obtain from the previous sequent by a single application of
$\forall E$).
\begin{proposition}\label{prop:qr} Let $\delta$ be first-order linear logic
proof in long normal form which has the translation of a hybrid sequent as its conclusion.
All
occurrences of $\forall E$ and $\forall I$ of the quantifier-reduced
from $\delta'$ of $\delta$ occur respectively
in the following contexts.
$$
\begin{array}{ccc}
\infer[\multimap E]{\Gamma,\Delta \vdash_n B}{\infer[\forall
E]{\Gamma\vdash_n A\multimap B}{\Gamma\vdash_n \forall x. [ A
\multimap B]} &\Delta \vdash_p A}
&&
\infer[\forall I]{\Gamma\vdash_p \forall x. [A \multimap
B]}{\infer[\multimap I]{\Gamma\vdash_p A\multimap
B}{\Gamma,A\vdash_p B}}
\end{array}
$$
\end{proposition}
\paragraph{Proof} Given Proposition~\ref{prop:forall} and the fact
that $\delta'$ is quantifier-reduced, all quantifiers occur in
(sub)formulas of the form $\forall x. [A \multimap B]$, which
corresponds to the translation of a Lambek formula. Given that
$\delta$ is in long normal form, meaning that the focus shift rule is
applied only to atomic formulas, so is its quantifier-reduced form $\delta'$.
Look at an arbitrary application of the $\forall E$ rule. We show it
must be part of a subproof of the form shown above on the left. After application of the
$\forall E$ rule, we have the sequent $\Gamma\vdash_n A\multimap
B$. The focus shift rule cannot apply, since $A\multimap B$ is not
atomic and $\delta'$ is in long normal form. Therefore, by inspection
of the available rules $\multimap E$ is the
only rule available and we are in the case shown above.
The case for the $\forall I$ rule is similar. To obtain a formula $\Gamma\vdash_p A\multimap
B$ as the premiss of the $\forall I$ rule, focus shift is excluded
because we have a complex formula. The only available alternative
removes the main connective as shown above on the right. \hfill \ensuremath{\Box}
\subsection{Hybrid proof to MILL1 proof}
After this long setup, everything is in place to prove the main
theorem. Thanks to the way we have defined our basic notions
and translations, the proof is rather simple. We show that under the
given translation, the proof rules of hybrid type-logical grammar are
derived rules of MILL1. In the next section, we show the converse:
that MILL1 proofs using formulas obtained from the translation
correspond to proofs in hybrid type-logical grammars.
The proof is actually stronger: we
show that proofs in the two systems generate the same
\emph{semantics}. This is easily verified since, as discussed in
Section~\ref{sec:sem},
the elimination (resp.\ introduction) rules for $/$, $\backslash$ and $|$ correspond to the
elimination (resp.\ introduction) rule for $\multimap$. The
elimination rules (for $/$, $\backslash$, $|$ and $\multimap$) correspond to application and the introduction rule
correspond to abstraction. The quantifier $\forall$ is treated as semantically inert.
\begin{lemma}\label{lem:htom} Let $\delta$ be a hybrid proof of $\Gamma \vdash A$,
then there is an MILL1 proof $\delta^*$ of its
translation $\| \Gamma \| \vdash \| A \|$.
\end{lemma}
\paragraph{Proof} We produce a unfocused proof with
unification (that is, we do not distinguish between $\vdash_p$ and
$\vdash_n$). If desired, we can transform the proof obtained by this
lemma into a focused proof by Proposition~\ref{prop:focus}). We also produce a proof in quantifier-reduced
form.
Since the lexicon/axioms rules are in beta-normal eta-long form by
definition, we know from Lemma~3.22 of \citeasnoun{kanazawa11pg} that substitution is only of type
variables/atoms for type variables and never of a complex type for a
type variable, so the arity of our predicate symbols in first-order
linear logic is fixed.
Induction on the depth $d$ of the proof.
If $d=1$ we either have an axiom rule or a lexical hypothesis. In both
cases, we have a sequent $x:A \vdash M^{\alpha}:A$, with $\alpha$ the
principal type of $M$ and with $x$ of type
$\alpha$ in the axiom case and of type $i \rightarrow i -1$ (for the $i$th
word) in the lexicon case.
We
translate the axiom by $\| A \|^{f(\alpha)}
\vdash \| A \|^{f(\alpha)}$ (letting the free variables of $f(\alpha)$
become free meta-variables)
and the lexical hypothesis by
the axiom $\| A \|_c^{f(\alpha)}
\vdash \| A \|^{f(\alpha)}$, replacing the meta-variables in
$f(\alpha)$ on the left with
variables and quantifying over them, making the formula on the
left-hand side of the turnstile closed. Since we produce a proof in
quantifier-reduced form, we do not perform the closure on the
right-hand side of the turnstile (or, if you prefer, we perform the
closure but immediately follow it by $\forall E$ rules for all
quantifiers introduced by the closure operation).
If $d >1$, induction hypothesis gives us proofs of the premisses of
the rule and we proceed by case analysis on the last rule in the
hybrid proof.
\begin{description}
\item{$[\backslash E]$} By induction hypothesis, we have a proof $\delta_1$ of
$\Gamma \vdash \|B \|^{[C,D]}$ and a proof $\delta_2$ of $\Delta\vdash \|
B\backslash A \|^{[F,E]}$. In addition, we know by induction hypothesis that
a MGU $\mathbb{s}$ of $\langle \Gamma;D\rangle$ and $\langle\Delta;F\rangle$ exists. Therefore, we can construct a proof of
the conclusion of the $/E$ rule as follows. Since $G$ is fresh,
unifying it with $C$ is possible and produces a new substitution $\mathbb{s}'$.
$$
\infer[\multimap E]{\mathbb{s}'(\Gamma),\mathbb{s}'(\Delta) \vdash \| A \|^{[\mathbb{s}'(C),\mathbb{s}'(E)]}}{
\infer*[\delta_1]{\Gamma \vdash \| B \|^{[C,D]}}{}
&
\infer[\forall E]{\Delta \vdash \| B \|^{[G,F]} \multimap \|
A \|^{[G,E]}}{\infer[=_{\textit{def}}]{\Delta \vdash \forall x. \| B \|^{[x,F]} \multimap \|
A \|^{[x,E]}}{\infer*[\delta_2]{\Delta \vdash \| B\backslash A \|^{[F,E]}}{}}}
}
$$
\item{$[/ E]$} Symmetric.
\item{$[| E]$} By induction hypothesis, we have a proof $\delta_1$ of
$\Gamma \vdash \| A| B \|^{f(\beta\rightarrow\alpha)}$ and a proof
$\delta_2$ of $\Delta \vdash \| B \|^{f(\gamma)}$. We also know
there is an MGU $\mathbb{s}$ of $\langle\Gamma;\beta\rangle$ and
$\langle\Delta;\gamma\rangle$. Therefore, we can combine these two
proofs using $\multimap E$ and (by Lemma~\ref{lem:mgu}) this same unification, as follows.
$$
\infer[\multimap E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta) \vdash \| A
\|^{f(\mathbb{s}(\alpha))}}{
\infer[=_{\textit{def}}]{\Gamma \vdash \| B \|^{f(\beta)} \multimap \| A
\|^{f(\alpha)}}{\infer*[\delta_1]{\Gamma \vdash \| A| B \|^{f(\beta\rightarrow\alpha)}}{}} &
\infer*[\delta_2]{\Delta \vdash \| B \|^{f(\gamma)}}{}
}
$$
\item{$[\backslash I]$} By induction hypothesis, we have a proof $\delta_1$
of $\Gamma, \| B \|^{[D,C]} \vdash \| A \|^{[D,E]}$. In addition,
since all principal types are balanced and the two occurrences of
$D$ occur in the translations of $B$ and $A$ respectively, we know
there are no occurrences of $D$ in $\Gamma$. Hence, after the
$\multimap I$ rule, we satisfy the
condition for the $\forall I$ rule and can extend the proof as follows.
$$
\infer[=_{\textit{def}}]{\Gamma \vdash \| B \backslash A \|^{[C,E]}}{
\infer[\forall I]{\Gamma \vdash \forall x. \| B \|^{[x,C]} \multimap \| A \|^{[x,E]} }{
\infer[\multimap I]{\Gamma \vdash \|B\|^{[D,C]} \multimap \| A \|^{[D,E]} }{
\infer*[\delta_1]{\Gamma, \| B \|^{[D,C]} \vdash \| A \|^{[D,E]} }{}
}
}}
$$
\item{$[/ I]$} Symmetric.
\item{$[| I]$} Induction hypothesis gives us a proof $\delta_1$ of
$\Gamma, \| B \|^{f(\beta)}\vdash \| A \|^{f(\alpha)}$, which we can
extend as follows.
$$
\infer[=_{\textit{def}}]{\Gamma \vdash \| A | B
\|^{f(\beta\rightarrow\alpha)}}{
\infer[\multimap I]{\Gamma \vdash \| B \|^{f(\beta)} \multimap \|A
\|^f(\alpha)}{
\infer*[\delta_1]{\Gamma, \| B \|^{f(\beta)}\vdash \| A \|^{f(\alpha)}}{}
}
}
$$
\hfill \ensuremath{\Box}
\end{description}
\subsection{MILL1 proof to hybrid proof}
\begin{lemma}\label{lem:mtoh} Let $\delta$ be the MILL1 derivation of the translation
of a hybrid sequent, that is, of
$\| A_1 \|^{[0,1]},\ldots, \| A_n \|^{[n-1,n]} \vdash \| B \|^{[0,n]}$. Then there is a
hybrid proof $\delta^*$ of $x_1^{1\rightarrow 0}:A_1,\ldots,x_n^{n\rightarrow
n-1}:A_n \vdash M^{n\rightarrow 0}:B$, where $M \equiv_{\beta\eta} \lambda
z. (x_1\, \ldots (x_n\, z))$.
\end{lemma}
\paragraph{Proof} The fact that $M \equiv_{\beta\eta} \lambda
z. (x_1\, \ldots (x_n\, z))$ follows immediately from the balanced
occurrences of the type constants $0, \ldots, n$.
Let $\delta$ be the focused MILL1 derivation of $\| \Gamma
\| \vdash_p \| A \|$, or, the case being, of $\| \Gamma
\| \vdash_n \| A \|$. We assume $\delta$ to be
in quantifier-reduced form.
We proceed by induction on the depth $d$ of the proof.
If $d=1$, then there are two cases.
\begin{description}
\item{\emph{Lexicon}} If the rule was a lexical
hypothesis, then it is a proof of $\| A_i \|^{[i-1,i]}$ for one of the
$A_i$ of the endsequent of the proof. By construction, we can recover the principal type $\alpha$ and (by
Coherence) a unique
$\beta$-normal $\eta$-long lambda term $M$ of type
$\alpha$. Therefore, we have a
hybrid proof $x_i^{i-1\rightarrow i}:A_i \vdash M^{\alpha}:A_i$, with $\alpha$
the principal type by construction.
\item{\emph{Axiom}} If the rule was an axiom then the formula $A$ does
not appear in the endsequent. We again recover the principal type
$\alpha$ and the
(eta-expanded) lambda term $M$ from the translation function and we return the hybrid proof
$x^{\alpha}:A \vdash M^{\alpha}: A$, with $M$ the eta-expansion of
$x$ to produce a valid Axiom rule.
\end{description}
If $d>1$, then we proceed by case analysis of the last rule of the
proof.
\begin{description}
\item{$[\pm]$} Induction hypothesis gives us the proof
corresponding to the negative premiss of the rule. We return the
same proof.
\item{$[\forall E/{}\multimap E]$} For the combination of a $\forall
E/\multimap E$ rule, there are two cases to consider, depending on
whether the translated formula had $/$ or $\backslash$ as main
connective. In case it was $/$, our translation unfolds as shown
below. The MGU
$\mathbb{s}$ unifies $D$ with $F$ (it doesn't matter here if the $\forall
E$ step has been done separately: in that case $x$ is replaced by a
fresh metavariable
$G$ and the MGU unifies $G$ with $E$).
$$
\infer[\multimap E]{\mathbb{s}(\| \Gamma\|), \mathbb{s}(\|\Delta \|) \vdash_n \| A \|^{[
\mathbb{s}(C),\mathbb{s}(E)]} }{\infer[\forall E]{\| \Gamma \| \vdash_n \| B
\|^{[D,E]}\multimap \| A \|^{[C,E]}}{\infer[=_{\textit{def}}]{ \| \Gamma \| \vdash_n
\forall x. \| B \|^{[ D,x ]} \multimap \| A \|^{[
C,x]}}{\| \Gamma \| \vdash_n \| A/B \|^{[C,D]}}} & \| \Delta \| \vdash_p \| B \|^{[F,E]}}
$$
Lemma~\ref{lem:mgu} guarantees that the two hybrid formulas $B$ are
indeed identical and
induction hypothesis gives us a proof $\delta_1$ of $\Gamma \vdash
M^{D\rightarrow C}:A/B$ and a proof $\delta_2$ of $\Delta \vdash N^{E\rightarrow F}:B$, which we can combine
by the $/ E$ rule, using the same substitution $\mathbb{s}$, to produce a
proof of $\Gamma,\Delta \vdash A^{\mathbb{s}(E)\rightarrow\mathbb{s}(C)}$ as required.
$$
\infer[/E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta)\vdash
(\lambda z.M (N\, z))^{\mathbb{s}(E)\rightarrow\mathbb{s}(C)}:A}{\infer*[\delta_1]{\Gamma \vdash M^{D\rightarrow
C}:A/B}{} & \infer*[\delta_2]{\Delta \vdash N^{E\rightarrow F}:B}{}}
$$
According to Lemma~\ref{lemma:pt}, we have also computed the corresponding
principal type $\mathbb{s}(E)\rightarrow\mathbb{s}(C)$.
The case for $\backslash$ is symmetric.
\item{$[\multimap E]$} In a quantifier reduced proof, a solitary
$\multimap E$ (without preceding $\forall E$ producing
the major premiss of the rule, which was treated in the previous case)
originated from a formula $A|B$. We
are in the following case.
$$
\infer[\multimap E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta) \vdash_n \| A
\|^{f(\mathbb{s}(\alpha))}}{
\infer[\equiv_{\textit{def}}]{\Gamma \vdash_n \| B \|^{f(\beta)} \multimap \| A \|^{f(\alpha)}}{
\Gamma \vdash_n \| A| B \|^{f(\beta\rightarrow\alpha)}} &
\Delta \vdash_p \| B \|^{f(\gamma)}}
$$
By
induction hypothesis there is a proof of $\delta_1$ of $\Gamma \vdash
A|B$ (where the term $M$ of $A|B$ has principal type $\beta\rightarrow \alpha$) and a proof $\delta_2$ of $\Delta
\vdash B$ (where the term $N$ assigned to $B$ has principal type
$\gamma$). By Lemma~\ref{lem:mgu}, the two hybrid $B$ formulas are
identical and we can use the MGU $\mathbb{s}$ as the most general unifier
of $\gamma$ and $\beta$. We can therefore combine these proofs
using the $| E$ rule and $\mathbb{s}$ as follows.
$$
\infer[|E]{\mathbb{s}(\Gamma),\mathbb{s}(\Delta)\vdash
(M\, N)^{\mathbb{s}(\alpha)}:A}{\infer*[\delta_1]{\Gamma\vdash M^{\beta\rightarrow\alpha}:A|B}{} & \infer*[\delta_2]{\Delta \vdash
N^{\gamma}:B}{}}
$$
Producing principal type $\mathbb{s}(\alpha)$ for this derivation.
\item{$[\multimap I/\forall I]$} If it results from a translation with
a pair of string formulas, we treat the combination of the
$\multimap I$ and a $\forall I$ rule as a single step. By Proposition~\ref{prop:qr}, we can do so without loss of
generality. Such a combination can only result from the
translation of a positive formula with main connective $/$ or $\I$. We treat only $/ I$; the
case for $\backslash I$ is symmetric.
$$
\infer[=_{\textit{def}}]{\Gamma \vdash_p \| A/ B
\|^{[C,D]}}{\infer[\forall I]{\Gamma
\vdash_p \forall x. \| B \|^{[D,x]} \multimap \| A \|^{[C,x]}}{
\infer[\multimap I]{\Gamma\vdash_p \| B \|^{[D,E]} \multimap \| A
\|^{[C,E]}}{
\Gamma,\|B \|^{[D,E]} \vdash_p \| A \|^{[C,E]}}
}
}
$$
We can simply extend the proof $\delta_1$ from the induction
hypothesis as follows.
$$
\infer[/I]{\Gamma \vdash ((\lambda x.M)(\lambda z.z))^{D\rightarrow C}:A/B}{\infer*[\delta_1]{\Gamma,
x^{E\rightarrow D}:B \vdash
M^{E\rightarrow C}:A}{}}
$$
\item{$[\multimap I]$} Finally, the case where the $\multimap I$ is
not followed by a $\forall I$ corresponds to the $| I$ rule. We are
in the following situation.
$$
\infer[=_{\textit{def}}]{\Gamma \vdash_p \| B| A
\|^{f(\beta\rightarrow\alpha)}}{\infer[\multimap I]{\Gamma\vdash_p \| B
\|^{f(\beta)} \multimap \| A \|^{f(\alpha)}}{
\Gamma, \| B \|^{f(\beta)} \vdash_p \| A \|^{f(\alpha)}}
}
$$
We can simply extend the proof $\delta_1$ of $\Gamma, x^{\beta}:B \vdash
M^{\alpha}:A$ given by the induction hypothesis as follows.
$$
\infer[| I]{\Gamma \vdash
(\lambda x.M)^{\beta\rightarrow\alpha}:A|B}{\infer*[\delta_1]{\Gamma ,x^{\beta}: B
\vdash M^{\alpha}:A}{}}
$$
\end{description}
\hfill \ensuremath{\Box}
\subsection{Main Theorem}
\label{sec:mainthm}
\begin{theorem}\label{thm:main} Derivability of hybrid type-logical grammars and their
translation into first-order linear logic coincides. Moreover,
proofs in the two systems produce the same semantic lambda terms.
\end{theorem}
\paragraph{Proof} Immediate from Lemma~\ref{lem:htom} and
Lemma~\ref{lem:mtoh} and the observation that $/E$, $\backslash E$ and $|E$, like
$\multimap E$ to which they correspond by translation,
are all translated as application on the meaning level and similarly for the different introduction rules and abstraction. \hfill \ensuremath{\Box}
Thanks to Theorem~\ref{thm:main}, we can use the well-understood proof
theory of first-order linear logic for parsing/theorem proving hybrid type-logical
grammars. Besides (focused) natural deduction and proof nets,
discussed in Section~\ref{sec:mill1}, the work on sequent proof search of
\citeasnoun{foll}, which includes a treatment of the additives, can
also directly be applied. These proof systems all have their strengths
and inconveniences, but, since they are all equivalent we can choose
the most appropriate tool for the job. For example, focused natural
deduction and proof nets
simplify the work of enumerating readings for a
given statement, and, as shown in Figure~\ref{fig:cycle}, proof nets provide an easy way to show
underivability of a statement.
In addition, the main theorem has the following immediate consequence.
\begin{corollary} Hybrid type-logical grammars are NP-complete
\end{corollary}
\paragraph{Proof} Hardness follows from the fact that hybrid
type-logical grammars contain the Lambek calculus (the implicational
fragment of the Lambek calculus was shown to be NP-complete by \citeasnoun{savateev}) --- or alternatively
from the fact that they contain lexicalized abstract categorial
grammars \cite{groote01acg}. Since first-order linear logic is NP-complete, by
Lemma~\ref{lem:htom} and the fact that the translation is linear in
the size of the formulas, hybrid type-logical grammars are in NP. \hfill \ensuremath{\Box}
To compare hybrid type-logical grammars with lambda grammars, we first
define an interesting subclass of hybrid type-logical grammars which we will show
to be equivalent to lambda grammars.
\begin{definition} A hybrid proof is \emph{strictly separated} iff for
every $/I$ and $\backslash I$ rule, the
subproof leading to the premiss of this introduction rule consists
only of Lambek elimination rules and premisses $A\vdash A$ with $A$
a Lambek formula (ie.\ a member of $\mathcal{F}_1$, containing only
$/$, $\backslash$ and simple atomic formulas).
\end{definition}
We can enforce strict separation directly in the proof theory by splitting the $\vdash$ symbol into
$\vdash_L$ and $\vdash_{\lambda}$, subscripting by $\vdash_L$ the premisses and
conclusions of the $/ E$, $\backslash E$, $/ I$, $\backslash I$ and
axiom/hypothesis for Lambek formulas as $\vdash_L$, subscripting by
$\vdash_{\lambda}$ the $| E$, $| I$ and axiom/hypothesis for formulas
not in $\mathcal{F}_1$ and adding the inclusion rule.
$$
\infer[L,\lambda]{\Gamma \vdash_{\lambda} M^{E\rightarrow
D}:B}{\Gamma\vdash_L M^{E\rightarrow D}:B}
$$
Not all proofs in hybrid type-logical grammars are strictly separated,
as shown by the example in Section~\ref{sec:hex} on page~\pageref{hexproof}, where the final $/I$
rule is preceded by both $|E$ and $|I$.
\begin{lemma}\label{lem:hybridlambda} Strictly separated hybrid type-logical grammars generate the same
string languages and the same string-meaning relations as lambda grammars.
\end{lemma}
\paragraph{Proof (sketch)} The main idea from \cite{busz96}, who uses a variant of
the proof from \cite{pentus,pentus97}, is that we can replace Lambek
calculus formulas by \emph{sets} of atomic formulas (CFG nonterminals)
which behave combinatorially like AB formulas --- the CFG nonterminals
are essentially the names for AB formulas --- in such a way that these sets
generate the same lambda term
semantics. Here, we do the same for all Lambek \emph{sub}-formulas of
a given hybrid type-logical grammar.
By the definition of strict separation, we know that all Lambek rules
occur in subproofs where these rules are not intermingled with the
lambda grammar rules. Hence, Buszkowski's construction translates
these proofs of $\Gamma \vdash_L B$ into proofs of $\Gamma' \vdash_L
B$ where only the $/E$ and $\backslash E$ rules are used. Then, by
treating all Lambek formulas as CFG nonterminals and all
instantiations of the $/E$ and $\backslash E$ rules in the grammar as
CFG rules. That is, the instantiation of the the $\backslash E$ rule for
specific formulas $A$ and $B$
$$
\infer[\backslash E]{B}{A & A\backslash B}
$$
\noindent becomes a non-logical rule
$$
\infer{C}{D & E}
$$
(or, if we prefer
to write it as a CFG rule: $D, E \longrightarrow C$), where $C$ is the non-terminal
corresponding to formula $B$, $D$ corresponds to formula $A$ and $E$
corresponds to the formula $A\backslash B$.
\editout{
Let $\mathcal{F}_1$ be the set of Lambek calculus formulas occurring as
subformulas of the lexicon of a hybrid type-logical grammar, let
$\mathcal{P}$ be the formulas of $\mathcal{F}_1$ which occur
positively (plus the atomic formula $s$) and $\mathcal{N}$ the
formulas of $\mathcal{F}_1$ which occur negatively. The sets
$\mathcal{N}$ and $\mathcal{P}$ exhaust the Lambek formulas which can
appear. By application of
the main theorem of \cite{busz96}, we produce an AB grammar
$G$ such that whenever $A_1, \ldots,
A_n \vdash C$ is derivable in the Lambek calculus (for all $A_i \in
\mathcal{N}$ and $C \in \mathcal{P}$), then the corresponding AB
grammar has formulas $B_i$
for each $A_i$ and $D$ for $C$ such that
$B_1,\ldots,B_n \vdash D$ is derivable using elimination rules only
\emph{and} this derivation produces the same semantics as the original proof.}
\hfill \ensuremath{\Box}
\begin{lemma}\label{lem:nplambda} Parsing lambda grammars which are
the translation of strictly separated hybrid type-logical grammars
is NP-complete.
\end{lemma}
\paragraph{Proof} The construction of Lemma~\ref{lem:hybridlambda}
generates, by means of the \citeasnoun{busz96} proof, many non-logical grammar
rules. Given that such a system may not be decidable, we need to
be careful. However, by the construction of \cite{busz96}, all
non-lexicalized rules are of the form $D, E \longrightarrow F$ with $D$, $E$ and
$F$ atomic formulas. Moreover, these atomic formulas correspond to AB
formulas, such that either $D = A/B$, $E=B$ and $F=A$ or $D=B$,
$F=B\backslash A$ and $F=A$ (for some Lambek formulas $A$ and
$B$). Therefore, we can start our proof by computing the closure of
these AB subproofs in $O(n^3)$, then continue the normal
lambda grammar proof, which is NP-complete. \hfill \ensuremath{\Box}
It should be obvious from the proof sketch of Lemmas~\ref{lem:hybridlambda} and~\ref{lem:nplambda} that though strictly
separated hybrid
type-logical grammars generate the same string languages and
string-meaning pairs as lambda grammars, hybrid type-logical grammars
allow a \emph{much} more compact specification of such grammars since
we avoid a brute-force explosion of the size of the lexicon and of the number of lexical entries per
word. Though I don't believe that the NP-complete problems we
encounter in computational linguistics are necessarily intractable ---
\citeasnoun{efficienthpsg} show that some NP-complete problems in
computational linguistics can be solved \emph{much} more
efficiently than $O(n^6)$ problems --- having an exponential explosion
of grammar size \emph{followed} by an NP-complete problem is
profoundly worrying for those interested in actually parsing the formalism.
It is unclear whether we can generalize the proof of Lemma~\ref{lem:hybridlambda} to dispense
with the strict separation requirement on hybrid grammars. Allowing
interleaving of the Lambek grammar and lambda grammar rules seems to
require a generalization of the results of \cite{busz96} to the hybrid
type-logical grammar case and, unless we change the proof of the
theorem considerably, this would require a type of
interpolation proof for hybrid type-logical grammars, which, as we will
see in Section~\ref{visual}, seems problematic for the lambda grammar
part of the system. For example, looking back to the proof in
Section~\ref{sec:hex}, it is unclear how to replace the final $/I$
rule by the elimination rule for either $/$ or $\backslash$, besides
adding $s/(np\backslash s)$ directly as an additional lexical entry
for the quantifier.
Also, though
it is certainly a desirable property of the hybrid system to derive
$s|(s|np) \vdash s/(np\backslash s)$ (for the given lexical lambda
term), since it relates the generalized quantifier formulas to one of
its standard Lambek calculus formulas, it is unclear if we actually
need this type of derivation to give a natural account of the
linguistic data. So the following question remains open: are there any
examples of hybrid type-logical grammar analyses where there is no
corresponding lambda grammar analysis? Having to resort to lexical
duplication is already a problem, both from a conceptual point of view
and from the point of view of parsing, but are there cases where even
this doesn't suffice?
Though we will leave this question unresolved, we investigate the descriptive
inadequacy of lambda grammars in Section~\ref{inadeq}.
\editout{
\section{Illustration}
In this section, I will give some examples of the translations of
abstract categorial grammar derivations into first-order linear logic.
The subscript $d$ for ``deep''. $m$, $l$ and $s$ should be seen as type
\emph{variables} (eg.\ $x_1$, $x_2$, $x_3$ etc.\ as used in typing
contexts $\Gamma$) and therefore different occurrences of the same
word will have distinct deep variables. The current choice (using the
initial letter of the word as much as possible) is
intended to make the correspondance with the words more clear.
$$
\infer[| E]{(s_d\ \lambda x. ((l_d\ x)\ m_d)):s}{\infer[| I]{\lambda x. ((l_d\ x)\ m_d):s|np}{\infer[| E]{((l_d\ x)\ m_d):s}{m_d:np & \infer[| E]{(l_d\ x):s|np}{l_d:(s|np)|np & [ x:np ]^1}}} & s_d:s|(s|np) }
$$
The lexicon $\rule{0pt}{1ex}^*$ translates the deep structure
constants into surface structure terms, which contain the surface
structure constants corresponding the the given word (that is, the
lexical term for $m_d$ contains the surface structure constant
$m_s$). These surface structure constants are of type string
($\sigma\rightarrow \sigma$), whereas the deep structure constants
correspond to the type given by the logical formula. So $(s|np)|np$
translates as a constant of type $(\sigma\rightarrow \sigma)\rightarrow
(\sigma\rightarrow\sigma) \rightarrow \sigma\rightarrow \sigma$.
{\renewcommand{\arraystretch}{1.3}
$$
\begin{array}{rl}
m_d^* &\vdash \lambda x^\sigma. (m_s^{\sigma\rightarrow \sigma}\ x) \\
l_d^* &\vdash \lambda O^{\sigma\rightarrow \sigma} \lambda S^{\sigma\rightarrow \sigma} \lambda y^\sigma. (S\ (l_s^{\sigma\rightarrow \sigma}\ (O\ y))) \\
s_d^* &\vdash \lambda P^{(\sigma\rightarrow \sigma) \rightarrow \sigma \rightarrow \sigma} \lambda z^\sigma. ((P\ s_s^{\sigma\rightarrow \sigma})\ z) \\
\end{array}
$$
}
We compute the principal type of each of the lexical entries, using a fresh variable for each. These principal types are all linear (sometimes called linear, non-deleting), therefore by the coherence theorem the principal type uniquely determines a $\lambda$-term (up to $\alpha\beta\eta$ equivalence). So given a principal type, the $\lambda$-term is superfluous.
{\renewcommand{\arraystretch}{1.3}
$$
\begin{array}{rl}
m_s^{1\rightarrow 0} &\vdash \lambda x^1. (m_s^{1\rightarrow 0}\ x):1\rightarrow 0 \\
l_s^{2\rightarrow 1} &\vdash \lambda O^{B\rightarrow 2} \lambda S^{1\rightarrow A} \lambda y^B. (S\ (l_s^{2\rightarrow 1}\ (O\ y))): (B\rightarrow 2) \rightarrow (1 \rightarrow A) \rightarrow B \rightarrow A \\
s_s^{3\rightarrow 2} &\vdash \lambda P^{(3\rightarrow 2) \rightarrow D \rightarrow C} \lambda z^D. ((P\ s_s^{3\rightarrow 2})\ z):((3\rightarrow 2) \rightarrow D \rightarrow C)\rightarrow D \rightarrow C \\
\end{array}
$$
}
Again, $m$, $l$ and $s$ should be seen as type variables (with the
subscript $s$ indicating ``surface'').
For a given sentence
$w_1^{0\rightarrow 1},\ldots,w_n^{n\rightarrow n-1}$ we want to construct a term $\tau$ of
type $n\rightarrow 0$. Given that this is a principal type, $\tau =_{\beta\eta}
\lambda x^{n} w_1 ( \ldots (w_n x))$, which is simply the lambda term
representation of the string $w_1,\ldots,w_n$. For the example above,
this means $\lambda x_3 . m_s (l_s (s_s x))$ or type $3\ra0$ in typing
context $\Gamma = m_s^{1\rightarrow 0}, l_s^{2\rightarrow 1}, s_s^{3\rightarrow 2}$.
As shown below, given the principal types of the lexicon, the
principal type inference algorithm allows us to compute the principal
type of each subproof. We use Hindley's algorithm \cite{hindley}, but with explicit type substitutions: in Hindley's presentation type substitution (or the computation of a MGU for two types) is integrated in the [$\multimap E$] rule, conflating the [$:=$] and [$\multimap E$] rules (ie.\ the explicit substitution is incorporated in the [$\multimap E$] rule directly below it).
The proof, using on-the-fly $\beta$-conversion to ensure all displayed
terms are $\beta$-normal $\eta$-long, is shown below. Type matching
(or a proof net presentation of the natural deduction proof) allows us
to infer the type of the hypothetical $x$. Also, since we only
consider $\eta$-long terms, substitution is only of type
variables/atoms for type variables and never of a complex type for a
type variable (Lemma 3.22 from \citeasnoun{kanazawa11pg}).
$$
\scalebox{.48}{
\infer[\multimap E]{m_s^{1\rightarrow 0}, l_s^{2\rightarrow 1}, s_s^{3\rightarrow 2} \vdash
\lambda y. (m_s\ (l_s (s_s\ y))): 3\rightarrow 0}{\infer[B:=3]{m_s^{1\rightarrow 0}, l_s^{2\rightarrow 1}\vdash \lambda x. \lambda y. (m_s\
(l_s\ (x\ y))):(3\rightarrow 2) \rightarrow 3 \rightarrow 0}{\infer[\multimap
I^1]{m_s^{1\rightarrow 0}, l_s^{2\rightarrow 1}\vdash \lambda x. \lambda y. (m_s\
(l_s\ (x\ y))):(B\rightarrow 2) \rightarrow B \rightarrow 0}{\infer[\multimap E]{m_s^{1\rightarrow
0}, l_s^{2\rightarrow 1},x^{B \rightarrow 2}\vdash \lambda y. (m_s\ (l_s\ (x\
y))): B \rightarrow 0 }{m_s^{1\rightarrow 0} \vdash \lambda x^1 (m_s\
x):\overset{np}{\overbrace{1\rightarrow 0}} & \infer[\multimap
E]{\infer[A:= 0]{l_s^{2\rightarrow 1},x^{B \rightarrow 2}\vdash\lambda S \lambda
y. (S\ (l_s\ (x\ y))): (1 \rightarrow 0) \rightarrow B \rightarrow 0}{l_s^{2\rightarrow
1},x^{B \rightarrow 2}\vdash\lambda S \lambda y. (S\ (l_s\ (x\
y))): (1 \rightarrow A) \rightarrow B \rightarrow A}}{l_s^{2\rightarrow 1} \vdash
\lambda O \lambda S \lambda y. (S\ (l_s\ (O\ y))):
\overset{np}{\overbrace{(B\rightarrow 2)}} \rightarrow
\overset{np}{\overbrace{(1 \rightarrow A)}} \rightarrow
\overset{s}{\overbrace{B \rightarrow A}} & \infer[F:=B]{x^{B\rightarrow
2}\vdash x:B\rightarrow 2}{\infer[E:=2]{x^{F \rightarrow 2} \vdash x:F\rightarrow 2 }{ [x^{F \rightarrow E} \vdash x:\overset{np}{\overbrace{F\rightarrow E}} ]^1}}}}}} & \infer[D:=3]{s_s^{3\rightarrow 2} \vdash \lambda P \lambda z. ((P\ s_s)\ z):((3\rightarrow 2) \rightarrow 3 \rightarrow 0)\rightarrow 3 \rightarrow 0}{\infer[C:=0]{s_s^{3\rightarrow 2} \vdash \lambda P \lambda z. ((P\ s_s)\ z):((3\rightarrow 2) \rightarrow D \rightarrow 0)\rightarrow D \rightarrow 0}{s_s^{3\rightarrow 2} \vdash \lambda P \lambda z. ((P\ s_s)\ z):(\overset{np}{\overbrace{(3\rightarrow 2)}} \rightarrow \overset{s}{\overbrace{D \rightarrow C}})\rightarrow \overset{s}{\overbrace{D \rightarrow C}}}} }}
$$
$$
\scalebox{.70}{
\infer[\multimap E]{1\rightarrow 0, 2\rightarrow 1, 3\rightarrow 2 \vdash 3\rightarrow 0}{
\infer[B:=3]{1\rightarrow 0, 2\rightarrow 1\vdash (3\rightarrow 2) \rightarrow 3 \rightarrow 0}{\infer[\multimap I^1]{1\rightarrow 0, 2\rightarrow 1\vdash (B\rightarrow 2) \rightarrow B \rightarrow 0}{
\infer[\multimap E]{1\rightarrow 0, 2\rightarrow 1, B \rightarrow 2 \vdash B \rightarrow 0 }{
1\rightarrow 0 \vdash 1\rightarrow 0
& \infer[\multimap E]{
\infer[A:= 0]{2\rightarrow 1,B \rightarrow 2\vdash (1 \rightarrow 0) \rightarrow B \rightarrow 0}{2\rightarrow 1,B \rightarrow 2\vdash (1 \rightarrow A) \rightarrow B \rightarrow A}}{
2\rightarrow 1 \vdash (B\rightarrow 2) \rightarrow (1 \rightarrow
A) \rightarrow B \rightarrow A & \infer[F:=B]{B\rightarrow 2\vdash
B\rightarrow 2}{\infer[E:=2]{F\rightarrow 2\vdash
F\rightarrow 2}{[ F \rightarrow E \vdash F\rightarrow E ]^1}}}}}} & \infer[D:=3]{3\rightarrow 2 \vdash ((3\rightarrow 2) \rightarrow 3 \rightarrow 0)\rightarrow 3 \rightarrow 0}{\infer[C:=0]{3\rightarrow 2 \vdash ((3\rightarrow 2)
\rightarrow D \rightarrow 0)\rightarrow D \rightarrow 0}{3\rightarrow 2 \vdash ((3\rightarrow 2) \rightarrow D
\rightarrow C)\rightarrow D \rightarrow C}}}
}
$$
Given that we can reconstruct the $\lambda$-terms from the antecedents
and the principal typing, we can therefore simplify the lexicon,
integrating all typing information directly in the formula
(translating $np:3\rightarrow 2$ as $np(2,3)$ to take the left-right order
into account; in general we can recover the complex structure of a
type from this list of arguments given that). Adding universal
quantifiers for all type variables not bound by the antecedent (and
pushing the quantifiers in as far as possible, ie. simplifying
$\forall X. A\multimap B$, with no occurrences of $X$ in $A$ by $A
\multimap\forall X. B$\footnote{If desired, we can also simplify $\forall
X. A \multimap B$, with no occurrences of $X$ in $B$ by $(\exists X.A)
\multimap B$ --- though if we do so we must be careful, since $\exists
X. (A\multimap B)$ does \emph{not} simplify further even when $X$
does not occur in $A$. The choice is inessential. However, it should be noted that $\forall$ can only occur in negative (sub-)formulas and $\exists$ only in positive (sub-)formulas.}) gives the following lexicon, in the $\forall, \multimap$ fragment of first-order linear logic.
$$
\begin{array}{rl}
m_s^{1\rightarrow 0} &\vdash np(0,1) \\
l_s^{2\rightarrow 1} &\vdash \forall B. np(2,B) \multimap \forall A. np(A,1) \multimap s(A,B)\\
s_s^{3\rightarrow 2} &\vdash \forall C\forall D. (np(2,3)\multimap s(C,D))\multimap s(C,D) \\
\end{array}
$$
Given this, the proof which corresponds directly to the proof above is
shown below. NOTE: this proof \emph{does not} correspond directly
since the $np$ hypothesis starts its life as $np(2,3)$ instead of
$np(E,F)$ (which it should for coherence to work correctly in the more
general, eg.\ when $np(A,A)$ is a possibility), see below for a proof
which does use $np(E,F)$.
$$
\scalebox{.8}{
\infer[\multimap E]{s(0,3)}{\infer[\multimap I^1]{np(2,3) \multimap s(0,3)}{\infer[\multimap E]{s(0,3)}{np(0,1) & \infer[\multimap E]{\infer[\forall E]{np(0,1) \multimap s(0,3)}{\forall A. np(A,1) \multimap s(A,3)}}{\infer[\forall E]{np(2,3) \multimap \forall A. np(A,1) \multimap s(A,3)}{\forall B. np(2,B) \multimap \forall A. np(A,1) \multimap s(A,B)} & [ np(2,3) ]^1}}} & \infer[\forall E]{ (np(2,3)\multimap s(0,3))\multimap s(0,3) }{\infer[\forall E]{\forall D. (np(2,3)\multimap s(0,D))\multimap s(0,D)}{\forall C\forall D. (np(2,3)\multimap s(C,D))\multimap s(C,D)}} }}
$$
The substitutions have been replaced by [$\forall E$] (in general,
there can be [$\exists I$] rules if the polarity requires it, however
there is nothing corresponding to either [$\forall I$] or [$\exists
E$], which is a source of problems for $\lambda$-grammars).
In theorem provers, it is often more convenient to replace explicit
substitution for the [$\forall E$] (and the [$\exists I$]) rules by
unification. In proof nets, these unifications occur at the axiom
links, In natural deduction proofs, they occur at the elimination
rules (for the introduction rules, we can use the most general hypothesis). In this representation, we end up with the following
Prolog-like notation, where quantifiers are implicit (that is, the
negative $\forall$ and the positive $\exists$ quantifiers will be
implicit; the positive $\forall$ and the negative $\exists$
quantifiers will still be explicit).
$$
\begin{array}{rl}
m_s^{1\rightarrow 0} &\vdash np(0,1) \\
l_s^{2\rightarrow 1} &\vdash np(2,B) \multimap np(A,1) \multimap s(A,B)\\
s_s^{3\rightarrow 2} &\vdash (np(2,3)\multimap s(C,D))\multimap s(C,D) \\
\end{array}
$$
And the proof of above will be simplified as shown below. Note that we
can still read of the instantiations of the variables
($A=0$,$B=3$,$C=0$,$D=3$,$E=2$,$F=3$) from the proof.
$$
\scalebox{.8}{
\infer[\multimap E]{s(0,3)}{\infer[\multimap I^1]{np(2,B) \multimap s(0,B)}{\infer[\multimap E]{s(0,B)}{np(0,1) & \infer[\multimap E]{np(A,1) \multimap s(A,B)}{np(2,B) \multimap np(A,1) \multimap s(A,B) & [ np(E,F) ]^1}}} & (np(2,3)\multimap s(C,D))\multimap s(C,D)}}
$$
The same proof, but with all substitutions done.
$$
\scalebox{.8}{
\infer[\multimap E]{s(0,3)}{\infer[\multimap I^1]{np(2,3) \multimap s(0,3)}{\infer[\multimap E]{s(0,3)}{np(0,1) & \infer[\multimap E]{np(0,1) \multimap s(0,3)}{np(2,3) \multimap np(0,1) \multimap s(0,3) & [ np(2,3) ]^1}}} & (np(2,3)\multimap s(0,3))\multimap s(0,3)}}
$$
\subsection{Atomic formulas with multiple string positions}
$$
\begin{array}{rl}
n^* &= \sigma \rightarrow \sigma \\
np^* &= \sigma \rightarrow \sigma \\
s^* &= \sigma \rightarrow \sigma \\
\textit{inf}^* &= (\sigma \rightarrow \sigma) \rightarrow \sigma \rightarrow \sigma \\
(A \multimap B)^* &= A^* \rightarrow B^* \\
\end{array}
$$
$$
\begin{array}{rl}
j_d &= np \\
h_d &= np \\
n_d &= np \\
g_d &= \textit{inf} \multimap np \multimap np \multimap s \\
r_d &= np \multimap \textit{inf} \\
\end{array}
$$
The linear order of the atomic types and type variables in the corresponding atomic formula does not matter --- as long as we translate them consistently back and forth from position in the type tree and position as argument --- but for types of order two or smaller (essentially: strings with one or more holes, corresponding to the well-nested MCFLs) we can use the following translation which respects the linear order of the types in the
final string (provided the abstraction are done ``from left to right'' as well. In the general case, on the final line below, there are $n+1$ distinct variables $A_i$, each occurring once. Since each argument (if any) is a string and the result is a string, there is always an even number of string positions. The special cases for two and four string positions are spelled out in the array below.
$$
\begin{array}{rl}
p:B\rightarrow A &\leadsto p(A,B) \\
p:(C\rightarrow B) \rightarrow (D\rightarrow A) &\leadsto p(A,B,C,D)\\
p:(A_2 \rightarrow A_1) \rightarrow \ldots (A_{n-1} \rightarrow A_{n-2}) \rightarrow A_n \rightarrow A_0 &\leadsto p(A_0,A_1,A_2,\ldots,A_{n-2},A_{n-1},A_n)\\
\end{array}
$$
{\renewcommand{\arraystretch}{1.3}
$$
\begin{array}{rl}
j_s^{1\rightarrow 0} &\vdash \lambda x^1. (m_j^{1\rightarrow 0}\ x):1\rightarrow 0 \\
h_s^{2\rightarrow 1} &\vdash \lambda y^2. (m_h^{2\rightarrow 1}\ y):2\rightarrow 1 \\
n_s^{3\rightarrow 2} &\vdash \lambda z^3. (m_s^{3\rightarrow 2}\ z):3\rightarrow 2 \\
g_s^{4\rightarrow 3} &\vdash \lambda I^{(4\rightarrow 3) \rightarrow D \rightarrow C} \lambda O^{C\rightarrow B} \lambda S^{B \rightarrow A} \lambda x'^D (S\ (O\ ((I\ g)\ x'))):\\
&\quad \overset{\textit{inf}}{\overbrace{((4\rightarrow 3) \rightarrow D \rightarrow C)}} \rightarrow \overset{np}{\overbrace{( C \rightarrow B)}} \rightarrow \overset{np}{\overbrace{(B \rightarrow A)}} \rightarrow \overset{s}{\overbrace{D \rightarrow A}} \\
r_s^{5\rightarrow 4} &\vdash \lambda P^{F\rightarrow E} \lambda Q^{4\rightarrow F} \lambda y'^5 (P\ (Q\ (r\ y'))): \overset{np}{\overbrace{ (F\rightarrow E) }} \rightarrow \overset{\textit{inf}}{\overbrace{ (4 \rightarrow F) \rightarrow 5 \rightarrow E }}
\end{array}
$$
}
$$
\begin{array}{rl}
j_s^{1\rightarrow 0} &= np(0,1) \\
h_s^{2\rightarrow 1} &= np(1,2) \\
n_s^{3\rightarrow 2} &= np(2,3) \\
g_s^{4\rightarrow 3} &= \textit{inf}(C,3,4,D) \multimap np(B,C) \multimap np(A,B) \multimap s(A,D) \\
r_s^{5\rightarrow 4} &= np(E,F) \multimap \textit{inf}(E,F,4,5) \\
\end{array}
$$
\infer[\multimap E]{s(0,5)}{ np(0,1) & \infer[\multimap E]{np(A,1) \multimap s(A,5)}{np(1,2) & \infer[\multimap E]{np(B,2) \multimap np(A,B) \multimap s(A,5)}{\textit{inf}(C,3,4,D) \multimap np(B,C) \multimap np(A,B) \multimap s(A,D) & \infer[\multimap E]{\textit{inf}(2,3,4,5)}{np(2,3) & np(E,F) \multimap\textit{inf}(E,F,4,5)}}}}
\infer[\rightarrow E]{n_s^{3\rightarrow 2} , r_s^{5\rightarrow 4}\vdash \lambda
Q^{4\rightarrow 3}\lambda y'^5 n(Q(r\ y')):(4\rightarrow 3) \rightarrow
5\rightarrow 2 }{n_s^{3\rightarrow 2} \vdash \lambda z^3. (m_s^{3\rightarrow 2}\
z):3\rightarrow 2 & r_s^{5\rightarrow 4} \vdash \lambda P^{F\rightarrow E} \lambda Q^{4\rightarrow
F} \lambda y'^5 (P\ (Q\ (r\ y'))): \overset{np}{\overbrace{ (F\rightarrow
E) }} \rightarrow \overset{\textit{inf}}{\overbrace{ (4 \rightarrow F) \rightarrow 5 \rightarrow
E } } }
}
\section{Comparison}
\label{sec:comparison}
The proof nets discussed in Section~\ref{sec:pn} provide an insightful way to
compare the different calculi discussed in this article in terms of
their basic ``building blocks'', seen from the
point of view of first-order linear logic.
We need to be careful, since this comparison only gives
\emph{necessary} conditions to be in a certain fragment of first-order
linear logic, and as such, we can use it only as a diagnostic for
showing that possibilities are \emph{absent} from a logic. We can
directly use the different translation functions to give sufficient
conditions.
The conditions on the variables in the different fragments are also
absent from the visual representation. Nevertheless, we will see that
this comparison is insightful.
\subsection{A visual comparison of the different calculi}
\label{visual}
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (forallnc) {$.$};
\node (forallnp) [above=2em of forallnc] {$.$};
\draw (forallnc) -- (forallnp);
\node (forallpc) [right=6em of forallnc] {$.$};
\node (forallpp) [above=2em of forallpc] {$.$};
\draw [dotted] (forallpc) -- (forallpp);
\node (tmplnl) [left=1.25em of forallnp] {};
\node (alollin) [above=2.5em of tmplnl] {$\ .\ $};
\draw (forallnp) -- (alollin);
\node (tmplnr) [right=1.25em of forallnp] {};
\node (blollin) [above=2.5em of tmplnr] {$\ .\ $};
\draw (forallnp) -- (blollin);
\node (tmplpl) [left=1.25em of forallpp] {};
\node (alollip) [above=2.5em of tmplpl] {$\ .\ $};
\node (tmplpr) [right=1.25em of forallpp] {};
\node (blollip) [above=2.5em of tmplpr] {$\ .\ $};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path [clip] (forallpp.center) circle (2.5ex) [reverseclip];
\end{pgfinterruptboundingbox}
\draw [dotted] (forallpp.center) -- (blollip);
\draw [dotted] (forallpp.center) -- (alollip);
\end{scope}
\begin{scope}
\path [clip] (alollip) -- (forallpp.center) -- (blollip);
\draw (forallpp.center) circle (2.5ex);
\end{scope}
\node (tmp) [left=3.7em of alollip] {};
\node (lab) [above=1em of tmp]
{$\overbrace{\rule{4.3em}{0pt}}^{\text{AB-grammar}} $};
\node (tmp) [left=0em of alollip] {};
\end{tikzpicture}
\end{center}
\caption{Lambek grammar}
\label{fig:llinks}
\end{figure}
Figure~\ref{fig:llinks} shows the Lambek calculus connectives as
links for first-order linear logic proof nets.
Curry's \citeyear{Curry61} criticism of the Lambek calculus connectives, seen from
the current perspective, is that they combine subcategorization information (functor-argument structure)
and string operations. Though from a modern proof-theoretical point
of view \cite{focus} it is perfectly valid to combine multiple
positive and multiple negative rules into a single rule, separating
the two gives more freedom (that is, it allows us to express more
relations between the string positions and go beyond simple
concatenation --- the prefix and postfix of the Lambek calculus).
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (forallnc) {$.$};
\node (forallnp) [above=2em of forallnc] {$.$};
\draw (forallnc) -- (forallnp);
\node (forallpc) [right=6em of forallnc] {$.$};
\node (forallpp) [above=2em of forallpc] {$.$};
\draw [dotted] (forallpc) -- (forallpp);
\node (lollinc) [above=5em of forallnc] {$.$};
\node (tmplnl) [left=1.25em of lollinc] {};
\node (alollin) [above=2.5em of tmplnl] {$\ .\ $};
\draw (lollinc) -- (alollin);
\node (tmplnr) [right=1.25em of lollinc] {};
\node (blollin) [above=2.5em of tmplnr] {$\ .\ $};
\draw (lollinc) -- (blollin);
\node (lollipc) [above=5em of forallpc] {$.$};
\node (tmplpl) [left=1.25em of lollipc] {};
\node (alollip) [above=2.5em of tmplpl] {$\ .\ $};
\node (tmplpr) [right=1.25em of lollipc] {};
\node (blollip) [above=2.5em of tmplpr] {$\ .\ $};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path [clip] (lollipc.center) circle (2.5ex) [reverseclip];
\end{pgfinterruptboundingbox}
\draw [dotted] (lollipc.center) -- (blollip);
\draw [dotted] (lollipc.center) -- (alollip);
\end{scope}
\begin{scope}
\path [clip] (alollip) -- (lollipc.center) -- (blollip);
\draw (lollipc.center) circle (2.5ex);
\end{scope}
\node (tmp) [left=3.7em of alollip] {};
\node (lab) [above=1em of tmp]
{$\overbrace{\rule{4.3em}{0pt}}^{\text{positive}} $};
\node (tmp) [right=0.9em of alollip] {};
\node (lab) [above=1em of tmp]
{$\overbrace{\rule{4.3em}{0pt}}^{\text{negative}} $};
\node (tmp) [left=0em of alollip] {};
\begin{pgfinterruptboundingbox}
\node (tmp) [left=6.5em of lollinc] {};
\node (lbr) [above=-1.0em of tmp] {$\text{\scriptsize
functor/argument}\left\{ \rule{0pt}{2.1em}\right.$};
\node (tmpb) [below=4em of tmp] {};
\node (lbrb) [above=-1.8em of tmpb] {$\text{\scriptsize
\ \ \ \ string positions}\left\{ \rule{0pt}{1.7em}\right.$};
\end{pgfinterruptboundingbox}
\end{tikzpicture}
\end{center}
\caption{MILL1}
\label{fig:lll}
\end{figure}
As shown in Figure~\ref{fig:lll}, the first-order linear logic solution decomposes the Lambek
connectives into separate subcategorization and string position
components. In a sense, this decomposition answers Curry's critique in
a very simple way.
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (forallnc) {$.$};
\node (forallnp) [above=2em of forallnc] {$.$};
\draw (forallnc) -- (forallnp);
\node (forallpc) [right=6em of forallnc] {};
\node (lollinc) [above=5em of forallnc] {$.$};
\node (tmplnl) [left=1.25em of lollinc] {};
\node (alollin) [above=2.5em of tmplnl] {$\ .\ $};
\draw (lollinc) -- (alollin);
\node (tmplnr) [right=1.25em of lollinc] {};
\node (blollin) [above=2.5em of tmplnr] {$\ .\ $};
\draw (lollinc) -- (blollin);
\node (lollipc) [above=5em of forallpc] {$.$};
\node (tmplpl) [left=1.25em of lollipc] {};
\node (alollip) [above=2.5em of tmplpl] {$\ .\ $};
\node (tmplpr) [right=1.25em of lollipc] {};
\node (blollip) [above=2.5em of tmplpr] {$\ .\ $};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path [clip] (lollipc.center) circle (2.5ex) [reverseclip];
\end{pgfinterruptboundingbox}
\draw [dotted] (lollipc.center) -- (blollip);
\draw [dotted] (lollipc.center) -- (alollip);
\end{scope}
\begin{scope}
\path [clip] (alollip) -- (lollipc.center) -- (blollip);
\draw (lollipc.center) circle (2.5ex);
\end{scope}
\node (tmp) [left=3.55em of alollip] {};
\node (lab) [above=1em of tmp]
{$\overbrace{\rule{4.3em}{0pt}}^{\text{2nd-order}\
\lambda\text{-grammar}} $};
\node (tmp) [left=0em of alollip] {};
\end{tikzpicture}
\end{center}
\caption{Lambda grammars}
\label{fig:lgl}
\end{figure}
Curry's own solution is different and causes a loss of symmetry: as
Figure~\ref{fig:lgl} makes clear, the
positive universal link is missing! This
loss of symmetry is easy to miss in a unification-based presentation of
the logic where, in addition, the quantifiers occur only as an implicit prefix of
the formula. For a
logician/proof theorist, this is worrying since many classical results
and desirable properties of the system (restriction to atomic axioms,
cut elimination, interpolation\footnote{Interpolation, proved first
for the Lambek calculus in \cite{Roorda} is a key component of the
context-freeness proof for the Lambek calculus of
\citeasnoun{pentus97} and is likely to play a similar role in proofs
about the generative capacity of these alternative and extended systems.}) depend on this symmetry. However, it is also the cause of empirical inadequacy: positive $A/B$ and $B\backslash A$ can no longer
be represented, hence no satisfactory treatment of adverbs,
coordination, gapping etc.; we will elaborate this point in detail in
Section~\ref{inadeq}.
Another way to look at this is that lambda grammars require
all formulas to be expressed in prenex normal form --- something we
exploit in the translation function. However, since
we are using linear logic, not all formulas have a prenex normal form.
The following are all underivable (assuming no occurrences of $x$ in
$B$). Refer back to Figure~\ref{fig:cycle} to see why the first
statement is underivable.
\begin{align*}
(\forall x. A) \multimap B &\nvdash \exists x (A
\multimap B) \\
\exists x (A
\multimap B) & \nvdash (\forall x. A) \multimap B \\
B \multimap \exists x. A & \nvdash \exists
x. (B\multimap A) \\
\exists
x. (B\multimap A) & \nvdash B \multimap \exists x. A
\end{align*}
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (forallnc) {$.$};
\node (forallnp) [above=2em of forallnc] {$.$};
\draw (forallnc) -- (forallnp);
\node (forallpc) [right=6em of forallnc] {};
\node (lollinc) [above=5em of forallnc] {$.$};
\node (tmplnl) [left=1.25em of lollinc] {};
\node (alollin) [above=2.5em of tmplnl] {$\ .\ $};
\draw (lollinc) -- (alollin);
\node (tmplnr) [right=1.25em of lollinc] {};
\node (blollin) [above=2.5em of tmplnr] {$\ .\ $};
\draw (lollinc) -- (blollin);
\node (l1) [above=1em of blollin] {$\ \ \qquad\overbrace{\rule{11.2em}{0pt}}^{\lambda\text{-grammar}}$};
\node (lollipc) [above=5em of forallpc] {$.$};
\node (tmplpl) [left=1.25em of lollipc] {};
\node (alollip) [above=2.5em of tmplpl] {$\ .\ $};
\node (tmplpr) [right=1.25em of lollipc] {};
\node (blollip) [above=2.5em of tmplpr] {$\ .\ $};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path [clip] (lollipc.center) circle (2.5ex) [reverseclip];
\end{pgfinterruptboundingbox}
\draw [dotted] (lollipc.center) -- (blollip);
\draw [dotted] (lollipc.center) -- (alollip);
\end{scope}
\begin{scope}
\path [clip] (alollip) -- (lollipc.center) -- (blollip);
\draw (lollipc.center) circle (2.5ex);
\end{scope}
\node (tmp) [left=13em of forallnc] {};
\node (forallnc) [above=2em of tmp] {$.$};
\node (forallnp) [above=2em of forallnc] {$.$};
\draw (forallnc) -- (forallnp);
\node (tmplnl) [left=1.25em of forallnp] {};
\node (alollin) [above=2.5em of tmplnl] {$\ .\ $};
\draw (forallnp) -- (alollin);
\node (tmplnr) [right=1.25em of forallnp] {};
\node (blollin) [above=2.5em of tmplnr] {$\ .\ $};
\draw (forallnp) -- (blollin);
\node (tmp) [left=13em of forallpc] {};
\node (forallppb) [above=5em of tmp] {$.$};
\node (forallpcb) [below=2em of forallppb] {$.$};
\node (tmplplb) [left=1.25em of forallppb] {};
\node (alollipb) [above=2.5em of tmplplb] {$\ .\ $};
\node (tmplprb) [right=1.25em of forallppb] {};
\node (blollipb) [above=2.5em of tmplprb] {$\ .\ $};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path[invclip] (forallppb.center) circle (2.5ex);
\end{pgfinterruptboundingbox}
\draw [dotted] (forallppb.center) -- (blollipb);
\draw [dotted] (forallppb.center) -- (alollipb);
\end{scope}
\begin{scope}
\path [clip] (alollipb) -- (forallppb.center) -- (blollipb);
\draw (forallppb.center) circle (2.5ex);
\end{scope};
\draw[dotted] (forallpcb) -- (forallppb);
\node (l2) [above=1em of blollin] {$\ \
\qquad\overbrace{\rule{11.2em}{0pt}}^{\text{Lambek grammar}}$};
\node (tmp) [left=7em of alollip] {};
\end{tikzpicture}
\end{center}
\caption{Hybrid grammar}
\label{fig:hgl}
\end{figure}
The hybrid solution to this problem is shown in Figure~\ref{fig:hgl}: reintroduce the positive Lambek connectives
directly. There are now two ways of coding the negative Lambek
connectives. The resulting system is also greater than the sum of its parts, since
gapping, which has a satisfactory neither in Lambek grammars nor in
lambda grammars, can be elegantly treated in hybrid categorial grammar
\cite{kl12gap,kl13emp}.
Symmetry is still lost\footnote{Neither full logical symmetry nor
having the Lambek calculus as a subsystems is of course necessary to
have an empirically valid formal system, as shown, for example by
CCG \cite{steedman}. However it calls for further investigation as to
what exactly is absent from the system and if this absence is
important from a descriptive point of view. For lambda grammars, we will do this in detail in Section~\ref{inadeq}.}, but empirically the system seems
comparable to the Displacement calculus \cite{mvf11displacement}: the Displacement calculus has
the full symmetry absent from hybrid type-logical grammars. In spite
of this,
as we have seen at the end of
Section~\ref{sec:string}, in many cases, the analyses proposed for the two
formalisms basically agree, as is made especially clear by their
translation into MILL1.
The differences between the two systems seems to be that hybrid
type-logical grammars can, like lambda grammars, generate
non-well-nested string languages and that Displacement grammars (seen
from the point of view of hybrid type-logical grammars) allow
the Lambek connectives to outscope the discontinuous connectives.
Further analysis is necessary to decide which of these two systems has
the better empirical coverage.
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (forallnc) {};
\node (forallnp) [above=2em of forallnc] {$.$};
\draw (forallnc) -- (forallnp);
\begin{pgfinterruptboundingbox}
\node (lbr) [left=2em of forallnp] {$\text{\scriptsize Lambek grammar}\left\{ \rule{0pt}{3.6em}\right.$};
\end{pgfinterruptboundingbox}
\node (bottn) [below=4em of forallnp] {};
\node (botn) [below=2em of bottn] {$.$};
\draw (bottn) -- (botn);
\node (dots) [above=.2em of bottn] {$\ldots$};
\node (forallpc) [right=6em of forallnc] {};
\node (forallpp) [above=2em of forallpc] {$.$};
\draw [dotted] (forallpc) -- (forallpp);
\node (bottp) [below=4em of forallpp] {};
\node (botp) [below=2em of bottp] {$.$};
\draw[dotted] (bottp) -- (botp);
\node (dots) [above=.2em of bottp] {$\ldots$};
\node (tmplnl) [left=1.25em of forallnp] {};
\node (alollin) [above=2.5em of tmplnl] {$\ .\ $};
\draw (forallnp) -- (alollin);
\node (tmplnr) [right=1.25em of forallnp] {};
\node (blollin) [above=2.5em of tmplnr] {$\ .\ $};
\draw (forallnp) -- (blollin);
\node (tmplpl) [left=1.25em of forallpp] {};
\node (alollip) [above=2.5em of tmplpl] {$\ .\ $};
\node (tmplpr) [right=1.25em of forallpp] {};
\node (blollip) [above=2.5em of tmplpr] {$\ .\ $};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path [clip] (forallpp.center) circle (2.5ex) [reverseclip];
\end{pgfinterruptboundingbox}
\draw [dotted] (forallpp.center) -- (blollip);
\draw [dotted] (forallpp.center) -- (alollip);
\end{scope}
\begin{scope}
\path [clip] (alollip) -- (forallpp.center) -- (blollip);
\draw (forallpp.center) circle (2.5ex);
\end{scope}
\node (tmp) [left=0.0em of alollip] {};
\end{tikzpicture}
\end{center}
\caption{D grammars, binary}\
\label{fig:dl}
\end{figure}
D grammars \cite{mvf11displacement} have a different perspective,
which is shown in Figure~\ref{fig:dl}. Functor argument structure
and string positions are still joined, but a greater number of
combinations are possible (from 0 to $n$ quantifiers, for a small
value of $n$ determined by the grammar). Lambek grammars are now the restriction to
a single quantifier for each binary connective.
\begin{figure}
\begin{center}
\begin{tikzpicture}
\node (forallnc) {};
\node (forallnp) [above=2em of forallnc] {$.$};
\draw (forallnc) -- (forallnp);
\node (bottn) [below=4em of forallnp] {};
\node (botn) [below=2em of bottn] {$.$};
\draw (bottn) -- (botn);
\node (dots) [above=.2em of bottn] {$\ldots$};
\node (bottp) [below=4em of forallpp] {};
\node (botp) [below=2em of bottp] {$.$};
\draw[dotted] (bottp) -- (botp);
\node (dots) [above=.2em of bottp] {$\ldots$};
\node (forallpc) [right=6em of forallnc] {};
\node (forallpp) [above=2em of forallpc] {$.$};
\draw [dotted] (forallpc) -- (forallpp);
\node (tmplnl) [left=1.25em of forallnp] {};
\node (alollin) [above=2.5em of tmplnl] {$\ .\ $};
\draw (forallnp) -- (alollin);
\node (tmplnr) [right=1.25em of forallnp] {};
\node (blollin) [above=2.5em of tmplnr] {$\ .\ $};
\draw (forallnp) -- (blollin);
\node (tmplpl) [left=1.25em of forallpp] {};
\node (alollip) [above=2.5em of tmplpl] {$\ .\ $};
\node (tmplpr) [right=1.25em of forallpp] {};
\node (blollip) [above=2.5em of tmplpr] {$\ .\ $};
\begin{scope}
\begin{pgfinterruptboundingbox}
\path [clip] (forallpp.center) circle (2.5ex) [reverseclip];
\end{pgfinterruptboundingbox}
\draw [dotted] (forallpp.center) -- (blollip);
\draw [dotted] (forallpp.center) -- (alollip);
\end{scope}
\begin{scope}
\path [clip] (alollip) -- (forallpp.center) -- (blollip);
\draw (forallpp.center) circle (2.5ex);
\end{scope}
\node (tt) [right=6em of bottp] {$.$};
\node (tb) [below=2em of tt] {$.$};
\draw (tt) -- (tb);
\node (pt) [right=6em of tt] {$.$};
\node (pb) [below=2em of pt] {$.$};
\draw[dotted] (pt) -- (pb);
\node (tmp) [right=5.0em of alollip] {};
\end{tikzpicture}
\end{center}
\caption{D grammars}
\label{fig:dlu}
\end{figure}
D grammars enriched with bridge, left projection and right projection,
shown in Figure~\ref{fig:dlu},
permit combinations of string position/subcategorization which
are not of the same polarity. These uses are rather restricted
compared to the visually similar quantifier link of first-order
linear logic:
essentially, they enable us to require that a pair of positions spans the
empty string.
Summing up, first-order linear logic decomposes the connectives of
different grammatical frameworks --- the Lambek calculus, lambda
grammars, Hybrid Type-Logical Grammars and the Displacement calculus
--- in a natural way into its four types of links. This visual
comparison both highlights the differences between this calculi and
opens the way for a more detailed comparison of the descriptive
limitations of one calculus compared to another.
Given that it is a decomposition of connectives, the MILL1 translation
is slightly bigger in terms of the total number of connectives in the
lexical entries. However, the basic operation are simple and
well-understood and the first-order variables actually function as
powerful constraints during proof search. Thanks to the embedding
results of this paper and of \cite{moot13lambek}, we can import the large range of linguistic
phenomena treated by Displacement grammars and Hybrid Type-Logical
Grammar directly into MILL1.
From the point of view of first-order linear logic, the connectives of
the other calculi are synthetic connectives, combined connectives of
the same polarity. We can mix and match these synthetic connectives as we see
fit. We can also exploit the symmetry of
first-order linear logic and use lambda grammar lexical entries as
arguments, restoring the symmetry of lambda grammars (and of Hybrid
Type-Logical Grammars). In addition, we can add the product $\otimes$ and
quantifier $\exists$ to our calculus essentially for free. Moreover,
as discussed in \cite{mill1,moot13lambek} we can use the quantifiers
of first-order linear logic to give an account of agreement and
island constraints as well. So we can improve upon Displacement
grammar analyses by adding agreement and island constraints and
improve upon Hybrid Type-Logical Grammar analyses by adding symmetry,
agreement and island constraints, all with the same logical primitives.
\section{Descriptive Inadequacy of Lambda Grammars}
\label{inadeq}
As already alluded to in Section~\ref{visual}, the asymmetry of
lambda grammars is the cause of descriptive inadequacy. Researcher in
lambda grammars have been aware of problems with coordination at least
since \citeasnoun{muskens01lfg}, who briefly mentions an apparent
incompatibility between lambda grammars
and the categorial grammar treatment of coordination, but the problem
can be traced back to \cite{Curry61} where the analysis of the coordination ``both
\ldots and \ldots'' in \S 5-6 is problematic. Kubota \& Levine \citeyear{kl13coord,kl13emp} show how
catastrophic the predictions of lambda grammars are; we will repeat
their observations below while adding several additional troublesome cases. This
problem has been little noted and little discussed\footnote{At least in the
lambda grammar and abstract categorial grammar literature, the problem
is discussed in the context of linear grammar in \cite{worth14coord}.}.
Indeed, one can find several
claims in the literature which deny there is a problem: Muskens claims
elsewhere \cite{muskens03lambda} that ``Since word order is now completely encoded in the
phrase structure term, there is no longer any need for a
directionality of the calculus'' and that ``The availability of
syntactic $\lambda$-terms reins in the overgeneration of the
traditional undirected calculi.''. However, as we will show below,
using lambda terms to limit the overgeneration of undirected calculi
is only partially successful and it is exactly for this reason that
a satisfactory treatment of coordination has remained elusive. Worse, the problem of overgeneration is not
limited to coordination, but a problem with \emph{any} higher-order type of the
Lambek calculus. The standard higher-order lambda grammar treatments
for generalized quantifiers and for non-peripheral are the only cases we know of where
lambda grammars make the right predictions. But even here, the
lambda grammar analysis does not generalize: generalized quantifiers
can be see as instances of Moortgat's \citeyear{m96q} $q(A,B,C)$
operator, and the lambda grammar treatment only works when $B$ is atomic and therefore for
quantifiers, of type $q(np,s,s)$, but not
for reflexives, of type $q(np,np\backslash s,np\backslash s)$.
For non-peripheral extraction, the lambda grammar analysis again
presupposes the extracted element is an atomic formula and therefore
the treatment does not generalize to gapping (for more on gapping see Section~\ref{sec:problems}).
To give an idea of how widespread and serious the problems are, the
following is a non-exhaustive list of problems for lambda grammars.
\ex. \label{sent:adverb} John deliberately hit Mary. (adverbs)
\ex. John bought a sandwich and ran to the train. (VP coordination)
\ex. \label{sent:fish} John caught and ate a fish. (TV coordination)
\ex. \label{sent:coord} John likes both black and gray t-shirts. (adjective coordination, after Curry, 1961)
\ex. \label{sent:rnr} John loves but Mary hates Noam. (right-node raising)
\ex. John bought himself a present. (reflexives)
\ex. John studies logic and Charles, phonetics. (gapping)
\ex. John left before Mary did. (ellipsis)
\ex. \label{sent:end} John ate more donuts than Mary bought bagels. (comparative sub-deletion)
These problems range from the mundane to the more involved, but the
important point is that, taken together, these problems occur
very frequently and that \emph{all} cases listed above have a simple and elegant
treatment in the Displacement calculus \cite{mvf11displacement}, in Hybrid
type-logical grammars \cite{kl12gap,kl13dgap} and in multimodal
type-logical grammars \cite{cgellipsis,KurtoMM}. Sentence~\ref{sent:adverb} to
\ref{sent:rnr} are simply and correctly handled by Lambek grammars
and Sentence~\ref{sent:adverb} to \ref{sent:coord} even by AB grammars.
Let me be precise about what I mean by descriptive inadequacy in this
context, since some authors use the term with a slightly different
meaning. A theory suffers from descriptive inadequacy if it fails to
capture linguistic generalizations and instead has to resort to
enumerating the linguistic data. In a lexicalized formalism like
categorial grammars, this means we want to avoid multiplying the
number of lexical
entries for the words in our grammar as much as
possible\footnote{Maybe a more reasonable measure would prefer the \emph{sum} of
the size for all entries assigned to a word to be as small as possible, since a single entry
$A_1 \oplus \ldots \oplus A_n$ is not really simpler that $n$ distinct
entries $A_1,\ldots,A_n$.
It should also be noted as the size of our grammar
increases (in terms of the number of words and constructions it is
able to handle), so does the size of our lexicon. So this is a relative
measure rather than an absolute one.}. So in the context of the examples
above, we would like Sentence~\ref{sent:adverb} to use the same
lexical entries as the sentence ``John hit Mary'', with the lexical
assignment to ``deliberately'' being to only addition and we would
like Sentence~\ref{sent:fish} to use the same lexical
entries as the sentences ``John caught a fish'' and ``John ate a
fish'', with the lexical assignment to ``and'' being the only
difference. When I say that lambda grammars suffer from descriptive
inadequacy, this does not mean that they are fundamentally unable to
handle Sentences~\ref{sent:adverb} to \ref{sent:end}, since Lemma~\ref{lem:hybridlambda}
guarantees that they can (given that the phenomena listed above all
have strictly separated hybrid proofs). I mean that they cannot treat the sentences
above without introducing otherwise unmotivated additional lexical
entries --- in fact, not without an exponential blowup of the size of the
lexicon, as is clear from Lemma~\ref{lem:hybridlambda}.
In Section~\ref{solutions} we will discuss the consequences of these problems in detail,
as well as some possible modifications to
lambda grammars which may solve these problems, chiefly among those are
extensions to hybrid type-logical grammar and to first-order linear logic.
\subsection{Inhabitation machines}
To show the main results, we need some additional notions of the typed
lambda calculus. An inhabitation machine (see \cite{barendregt13types}) is a type of grammar which,
given a type, enumerates all possible terms of this type. Their use
for categorial grammars has been pioneered by \citeasnoun{benthem}.
From page 33 of \cite{barendregt13types}, the following two-level grammar
(defined on type-context pairs) enumerates all closed
inhabitants in beta-normal eta-long form of a given type.
$\Gamma$ is a context, $\Gamma,x^{\alpha}$ denotes $\Gamma\cup \{
x^{\alpha} \}$ (where $x$ is distinct from the terms in $\Gamma$, so the result is again a
valid context), $A$ is an atomic type, $\alpha$, $\beta$ are
arbitrary types and $\vec{\alpha} \rightarrow \beta$ is short for $\alpha_1 \rightarrow
\ldots \rightarrow \alpha_n \rightarrow \beta$.
$$
\begin{array}{rl}
L(A;\Gamma) &\Longrightarrow xL(\alpha_1;\Gamma)\ldots L(\alpha_n;\Gamma)
\text{\qquad\qquad if}\ x:\vec{\alpha}\rightarrow A \in \Gamma \\
L(\alpha\rightarrow\beta;\Gamma) &\Longrightarrow \lambda x^{\alpha}. L(\beta;\Gamma,x^{\alpha})
\end{array}
$$
The lambda grammar case is considerably more restricted: the lexical
lambda terms must be linear and contain, for a given
word $w$ with corresponding variable $w_s$, a single occurrence of $w_s$. That is, we start with
$\Gamma = \{ w_s^{\sigma\rightarrow \sigma} \}$ and for the application rule, we partition
$\Gamma - \{ x^{\vec{\alpha} \rightarrow A} \} $ into jointly exhaustive, pairwise disjoint subsets
and divide these over the different subterms. In addition, we want our
lexical term to produce the correct word order and to be compatible
with the syntactic lambda grammar derivation.
\subsection{Problems for lambda grammars}
\label{sec:problems}
In this next section, we will show several problematic cases for lambda
grammars, using inhabitation
machines to exhaust all possible solutions and find all of them inadequate.
\subsubsection*{Adverbs}
As a first problem for lambda grammars, the Lambek calculus formula of an
adverb such as ``deliberately'', as it occurs in a sentence like
``Eduardo deliberately fell'', is $(np\backslash s)/(np
\backslash s)$ --- it modifies a verb having taken all arguments
except its subject and this verb phrase is on the immediate right of
the adverb. If we translate this formula to a first-order formula and
move (where possible) the quantifiers to the prefix and eliminate
them, we obtain the formula $(\forall
c. np(c,2) \multimap s(c,D)) \multimap np(E,1) \multimap s(E,D)$ but
we cannot use the principal type $((2\rightarrow c) \rightarrow D \rightarrow c)\rightarrow (1\rightarrow E) \rightarrow D\rightarrow E$
(with $c$ a fresh type constant) since it is uninhabited.
The lambda grammar syntactic type $(s|np) | (s|np)$ translates to the
prosodic type $((\sigma\rightarrow\sigma) \rightarrow \sigma\rightarrow\sigma) \rightarrow (\sigma\rightarrow\sigma) \rightarrow \sigma\rightarrow\sigma$ and
produces the inhabitation machine shown in Figure~\ref{fig:inhab}. We use the variable $d$
(of type $\sigma\rightarrow\sigma$)
to stand for the occurrence of the string ``deliberately''. We can see that the $\textit{VP}$ node in
the figure requires first an argument of type $\sigma\rightarrow\sigma$ (the
downward arrow) then an argument of type $\sigma$ (the upward arrow)
to produce a term of type $\sigma$. Valid linear paths through the
machine must pass each term label exactly once, and must pass the
$\lambda y$-label (on the curved arrow upwards to $\sigma$) before the $y$ variable.
\begin{figure}
\begin{center}
\begin{tikzpicture}[node distance=1.5cm]
\node[draw] (top) {$((\sigma\rightarrow\sigma) \rightarrow \sigma\rightarrow\sigma) \rightarrow (\sigma\rightarrow\sigma) \rightarrow \sigma\rightarrow\sigma$};
\node[draw, below of=top] (sigma) {$\sigma$};
\path[->] (top) edge node[right] {$\lambda
\textit{VP}^{(\sigma\rightarrow\sigma)\rightarrow\sigma\rightarrow\sigma}
\textit{NP}^{\sigma\rightarrow\sigma} z^{\sigma}$} (sigma);
\path[->] (sigma) edge [loop left] node[left] {$d$} ()
edge [loop right] node[right] {$\textit{NP}$} ();
\node[below of=sigma] (z) {$z$};
\node[left of=z] (vp) {$VP$};
\node[right of=z] (y) {$y$};
\path[->] (sigma) edge node {} (vp);
\path[->] ([xshift=-1pt] vp.north east) edge node {} ([xshift=-3pt] sigma.south);
\path[->] (sigma) edge node {} (y);
\path[->] (sigma) edge node {} (z);
\node[draw,below of=vp] (sisi) {$\sigma\rightarrow\sigma$};
\path[->] (vp) edge node {} (sisi);
\draw[->] (sisi.north west) to[out=145,in=115] (sigma.north west);
\node[left of=vp,node distance=1.15cm] {$\lambda y^{\sigma}$};
\end{tikzpicture}
\end{center}
\caption{Inhabitation machine for an adverb type.}
\label{fig:inhab}
\end{figure}
Figure~\ref{fig:inhabb} spits the $\sigma$ node in two, making the scope of
the $y$ variable clearer.
\begin{figure}
\begin{center}
\begin{tikzpicture}[node distance=1.5cm]
\node[draw] (top) {$((\sigma\rightarrow\sigma) \rightarrow \sigma\rightarrow\sigma) \rightarrow (\sigma\rightarrow\sigma) \rightarrow \sigma\rightarrow\sigma$};
\node[draw, below of=top] (sigma) {$\sigma$};
\path[->] (top) edge node[right] {$\lambda
\textit{VP}^{(\sigma\rightarrow\sigma)\rightarrow\sigma\rightarrow\sigma}
\textit{NP}^{\sigma\rightarrow\sigma} z^{\sigma}$} (sigma);
\path[->] (sigma) edge [loop left] node[left] {$d$} ()
edge [loop right] node[right] {$\textit{NP}$} ();
\node[below of=sigma] (tmp) {};
\node[left of=tmp] (vp) {$VP$};
\node[right of=tmp] (z) {$z$};
\path[->] (sigma) edge node {} (vp);
\path[->] ([xshift=-1pt] vp.north east) edge node {} ([xshift=-3pt] sigma.south);
\path[->] (sigma) edge node {} (z);
\node[draw,below of=vp] (sisi) {$\sigma\rightarrow\sigma$};
\path[->] (vp) edge node {} (sisi);
\node[draw,below of=sisi] (sbis) {$\sigma$};
\path[->] (sisi) edge node[right] {$\lambda y^{\sigma}$} (sbis);
\node[below of=sbis] (y) {$y$};
\path[->] (sbis) edge node {} (y);
\path[->] (sbis) edge [loop left] node[left] {$d$} ()
edge [loop right] node[right] {$\textit{NP}$} ();
\end{tikzpicture}
\end{center}
\caption{Simplified inhabitation machine for an adverb type.}
\label{fig:inhabb}
\end{figure}
The word order of the sentence constrains the paths we can take. We
must take an $\textit{NP}$ arc before we take a $d$ arc, since
``deliberately'' occurs after the subjet noun phrase. So from the top
$\sigma$ node, we can only take three possible paths, as shown
below. For comparison, the uninhabited type corresponding most closely
to the first-order formula is shown as item~\ref{uninhabited}. We can
see that the three other types are obtained by replacing the $c$
constant by a $C$ variable and exchanging one of the occurrences of
$C$ with another atomic type in such a way that the resulting type is inhabited.
\begin{align}
\label{delibone} \lambda \textit{VP} \lambda \textit{NP} \lambda z. & \textit{NP}\ (d\
((\textit{VP}\ \lambda y.y)\ z)): \\ & ((C\rightarrow C)\rightarrow D\rightarrow 2) \rightarrow (1\rightarrow E)
\rightarrow D \rightarrow E \nonumber \\
\label{delibtwo} \lambda \textit{VP} \lambda \textit{NP} \lambda
z. &\textit{NP}\ ((\textit{VP}\ \lambda y.d\ y)\ z): \\
& ((2\rightarrow 1)\rightarrow D\rightarrow C) \rightarrow (C\rightarrow E)
\rightarrow D \rightarrow E \nonumber \\
\label{delibthree} \lambda \textit{VP} \lambda \textit{NP} \lambda z. & ((\textit{VP}\
\lambda y.\textit{NP}\ (d\ y))\ z): \\
& ((2\rightarrow C)\rightarrow D\rightarrow E) \rightarrow (1\rightarrow C)
\rightarrow D \rightarrow E \nonumber \\
\label{uninhabited} \textit{Uninhabited:} \\
& ((2\rightarrow c) \rightarrow D \rightarrow c)\rightarrow (1\rightarrow E) \rightarrow D\rightarrow E \nonumber
\end{align}
We investigate the three possibilities in turn.
Lambda term~\ref{delibone} comes closest to the first-order linear logic formula, but it
is a lambda term modeled after those used for extraction and, as such, it takes
a sentence missing a noun phrase \emph{anywhere} as its argument, instead of
a verb phrase. Therefore, it
incorrectly predicts that the three following sentences are all grammatical.
\ex. \label{wronga} John deliberately Mary hit.
\ex. \label{wrongb} John deliberately Mary insinuates likes
Susan.
\ex. \label{wrongc} John deliberately Mary hit the sister of.
Predicting that sentence~\ref{wronga} means
``It was deliberate on the part of John that Mary hit him'', with sentence~\ref{wrongb} meaning approximately ``John made Mary insinuate that he likes Susan''
and sentence~\ref{wrongc} meaning something like ``Mary hit the
sister of John and this was deliberate on the part of John''. It seems
very difficult to block this example without also blocking the noun
``boy which Mary likes the sister of'' (not super-natural, but we want
to allow these kinds of extractions which are essentially
indistinguishable from the current formula).
Lambda term~\ref{delibtwo} shifts from the extraction-like lambda term
and its corresponding overgeneration to a lambda term similar to those
used for in situ binding/quantifying in\footnote{As we have seen, a
generalized quantifier like ``everyone'' is assigned the lambda term
$\lambda P.P(e)$ with $e$ being the string constant corresponding to
the word ``everyone''.},
where we require a sentence missing a noun phrase at the position of
``deliberately'' as argument. Though this analysis again allows us to
derive the correct word order, it also makes
the dubious claim that there is an $np$ constituent at the position of the
adverb. In addition, it overgenerates as follows.
\ex. \label{wrongd} Mary John hit deliberately.
\ex. \label{wronge} Mary the friend of deliberately left.
\ex. \label{wrongf} Mary John gave the friend of deliberately a book.
Though it is possible to argue that sentence~\ref{wrongd} is a sort of
topicalization (with stress on \emph{Mary}), it is problematic that this topicalization is
triggered by the adverb, since topicalization is independent of the
presence or absence of adverbs. Moreover, we generate the semantics ``It was deliberate on the part of Mary that
John hit her'' for sentence~\ref{wrongd}. We generate the semantics ``It was deliberate on
the part of Mary that her friend left'' for sentence~\ref{wronge} and
similarly ``Mary incited John to give her friend a book'' for sentence~\ref{wrongf}.
Finally, lambda term~\ref{delibthree} selects for a sentence missing a
noun phrase with the only condition that this noun phrase occurs
directly before the adverb. Here, we make the odd claim that the noun
phrase and the adverb \emph{together} span the position of an $np$:
that is, it claims that an adverb is a post-modifier of an $np$. In
addition, it is again an in situ binding/quantifying in analysis, but this time with
the complex string ``$np$ deliberately'' (where lambda
term~\ref{delibtwo} used an in situ binding analysis of just the word ``deliberately'').
\ex. \label{wrongg} John hit Mary deliberately.
\ex. \label{wrongh} The friend of Mary deliberately left.
\ex. \label{wrongi} The friend of Mary deliberately who lives in Paris
left.
Though sentences~\ref{wrongg} and \ref{wrongh} are syntactically
correct, the problem is that we generate the semantics ``It was deliberate on the part of Mary that
John hit her'' for sentence~\ref{wrongg} and a reading ``It was
deliberate on the part of Mary that her friend left'' for
sentence~\ref{wrongh} and \ref{wrongi}.
In sum, we cannot capture the essence of the Lambek calculus
formula $(np\backslash s)/(np
\backslash s)$ in lambda grammars. Other adverb formulas --- $(np\backslash s)\backslash (np
\backslash s)$ (an adverb occurring after the verb phrase) and $(n/n)/(n
/ n)$ (for adverbs such as ``very''), etc.\ --- suffer from the same
problem. The best approximations
that we can obtain all suffer from overgeneration because
non-com\-mu\-ta\-ti\-vi\-ty is insufficiently enforced.
There is, of course, a solution which replaces the complex $np\backslash s$
argument by a new atomic formula, say $vp$ and then, for all
lexical items of the form $((np\backslash s)/A_n)\ldots /A_1$, adds an
additional formula $(vp/A_n)\ldots / A_1$. This would essentially double the
number of lexical formulas for verbs, adverbs and prepositions --- syntactic
categories which already have a high number of lexical formulas ---
for just a single type of problematic example... More such examples
will follow.
We
will discuss this potential solution in a bit more detail in
Section~\ref{solutions}, but it should already be clear that this is
not a particularly attractive option, since it is a prototypical
example of descriptive inadequacy, the reasons for doubling
the lexicon are purely theory-internal: no other categorial grammar,
not even AB grammars, have this type of overgeneration for the simple
cases we've shown.
\subsubsection*{Coordination}
As noted by Kubota \& Levine \citeyear{kl13coord,kl13emp}, we can play a similar game for ``John
caught and ate a fish'', which looks as shown in Figure~\ref{fig:inhabtv}; for the sake of
space, we do not show the prefix $\lambda \textit{TV2}. \lambda \textit{TV1}. \lambda \textit{NP2}.
\lambda \textit{NP1}. \lambda z$, where $\textit{TV2}$ is the
transitive verb to the right of ``and'' (``ate'' in the current
example), $\textit{TV1}$ is the transitive verb to the left of it
(``caught''), $\textit{NP1}$ is the subject, $\textit{NP2}$ is the
object and $z$ is the end of the complete string.
Remark that ``and'' takes all constituents as argument: the two
transitive verbs, the subject noun phrase and the object noun phrase,
so it would seem that we \emph{should} be able to generate the right string.
\begin{figure}
\begin{center}
\begin{tikzpicture}[node distance=1.5cm]
\node[draw] (SIGMA) {$\ \sigma\ $};
\path[->] (SIGMA) edge [loop above] node[above] {$\textit{and}$} ()
edge [in=30,out=50,loop] node[above] {$\ \ \ \ \textit{NP2}$} ()
edge [in=130,out=150,loop] node[above]
{$\textit{NP1}\ \ \ \ $} ();
\node (Z) [below of=SIGMA] {$z$};
\node (TV1) [left of=Z] {$\textit{TV1}$};
\node (TV2) [right of=Z] {$\textit{TV2}$};
\draw[->] (SIGMA.west) -- ([xshift=-3pt] TV1.north);
\draw[->] ([xshift=3pt] TV1.north) -- (SIGMA.south west);
\draw[->] (SIGMA.east) -- ([xshift=3pt] TV2.north);
\draw[->] ([xshift=-3pt] TV2.north) -- (SIGMA.south east);
\draw[->] (SIGMA.south) -- (Z.north);
\node[draw] (OBJ1) [below of=TV1] {$\sigma\rightarrow\sigma$};
\node[draw] (SUBJ1) [left of=OBJ1] {$\sigma\rightarrow\sigma$};
\node[draw] (SUBJ11) [below of=SUBJ1] {$\sigma$};
\node[draw] (OBJ11) [below of=OBJ1] {$\sigma$};
\node (X) [below of=SUBJ11] {$x$};
\node (Y) [below of=OBJ11] {$y$};
\path[->] (SUBJ11) edge [loop left] node[left] {$\textit{NP1}$} ();
\path[->] (OBJ11) edge [loop right] node[right] {$\textit{and}$} ();
\path[->] (SUBJ1) edge node[left] {$\lambda x$} (SUBJ11);
\path[->] (OBJ1) edge node[left] {$\lambda y$} (OBJ11);
\draw[->] (TV1) -- (OBJ1);
\draw[->] (TV1) -- (SUBJ1);
\draw[->] (OBJ11) -- (Y);
\draw[->] (SUBJ11) -- (X);
\node[draw] (SUBJ2) [below of=TV2] {$\sigma\rightarrow\sigma$};
\node[draw] (OBJ2) [right of=SUBJ2] {$\sigma\rightarrow\sigma$};
\node[draw] (SUBJ22) [below of=SUBJ2] {$\sigma$};
\node[draw] (OBJ22) [below of=OBJ2] {$\sigma$};
\node (V) [below of=SUBJ22] {$v$};
\node (W) [below of=OBJ22] {$w$};
\draw[->] (TV2) -- (OBJ2);
\draw[->] (TV2) -- (SUBJ2);
\draw[->] (OBJ22) -- (W);
\draw[->] (SUBJ22) -- (V);
\path[->] (SUBJ2) edge node[left] {$\lambda v$} (SUBJ22);
\path[->] (OBJ2) edge node[left] {$\lambda w$} (OBJ22);
\path[->] (SUBJ22) edge [loop left] node[left] {$\textit{and}$} ();
\path[->] (OBJ22) edge [loop right] node[right] {$\textit{NP2}$} ();
\end{tikzpicture}
\end{center}
\caption{Simplified inhabitation machine for transitive verb
conjunction}
\label{fig:inhabtv}
\end{figure}
As before, we have split the $\sigma$ and $\sigma\rightarrow\sigma$ nodes for
readability; the actual graph merges all $\sigma$ and all
$\sigma\rightarrow\sigma$ nodes. The implausible analyses with $\textit{TV1}$
and subject of $\textit{TV2}$ and with $\textit{TV2}$ as object of
$\textit{TV1}$ are not shown in the figure, but they fail for the same
reasons discussed below.
The graph of Figure~\ref{fig:inhabtv} shows that the $\textit{TV1}$ node takes first its subjet (down and to the left
of it), then its object (directly below) and finally an argument of
type $\sigma$ (the upward arrow back to $\sigma$) and similarly for
$\textit{TV2}$. The $\textit{TV1}$ node (optionally) takes
$\textit{NP1}$ as its subjet and $\textit{TV2}$ (optionally) takes
$\textit{NP2}$ as its object.
Two combinations are fairly limited: the second argument of
$\textit{TV1}$ is either $\textit{NP1}$ or the empty string and the
first argument of $\textit{TV2}$ is either $\textit{NP2}$ or the empty
string. However, if the lexical entry contains the subterm
$((\textit{TV1}\, M)\, \textit{NP1})$ (for some $M$ at the place of
the object), then we are
essentially using a quantifying-in analysis for the subject: $(\textit{TV1}\, M)$ is a
sentence missing a noun phrase \emph{anywhere} and applying this term
to an argument puts this argument back at the place of the missing
noun phrase. Consequently, it would allow the derivation of ``caught
John and ate a fish''. Similar overgeneration occurs for ``ate'' and
``a fish'' if we use the quantifying-in combination $(\textit{TV2}\,
\textit{NP2})$ for the object.
If we want to avoid both types of overgeneration (subject
quantifying-in and object quantifying in),
the only remaining analysis
consists of choosing $\lambda x.x$,
$\lambda y.y$, $\lambda v.v$ and $\lambda w.w$ as arguments for the
two transitive verbs.\footnote{This solution still overgenerates
because it equates transitive verb with ``sentence missing two noun
phrases'' and therefore incorrectly predicts that ``John likes $[]_{np}$ 's
friend from $[]_{np}$'' can felicitously fill this role as follows.
\ex. Mary went to and John likes 's friend from Paris.
Meaning ``Mary went to Paris and John likes Mary's friend from there''.}
This solution is shown in full below.
\begin{align*}
\lambda \textit{TV2}. &\lambda \textit{TV1}. \lambda \textit{NP2}.
\lambda \textit{NP1}. \lambda z. \\ &\textit{NP1}\, ((\textit{TV1}\, \lambda
x. x\, \lambda y.y) (\textit{and}\, ((\textit{TV2}\, \lambda
v. v\, \lambda w.w) (\textit{NP2}\, z))))
\end{align*}
\editout{
$$
\infer{s}{\infer{np}{\textit{John}} &
\infer{np\rightarrow s}{
\infer{np\rightarrow np\rightarrow s}{
\infer{np\rightarrow np\rightarrow s}{\textit{caught}} &
\infer{a2}{\infer{a1}{\textit{and}} & \infer{np\rightarrow
np\rightarrow s}{\textit{ate}}}}
& \infer{np}{\textit{a fish}}}}
$$
}
\begin{figure}
\begin{center}
\begin{sideways}
$$
\infer{\overset{((((\textit{and}\,
a)\, c)\, f)\, j)^{5\rightarrow 0}}{s}}{\infer{np^{1\rightarrow 0}}{\textit{John}} &
\infer{\overset{(((\textit{and}\,
a)\, c)\, f)^{(1\rightarrow
K)\rightarrow 5\rightarrow K}}{s|np}}{
\infer{\overset{((\textit{and}\,
a)\, c)^{(L\rightarrow 4)\rightarrow (1\rightarrow
K)\rightarrow L\rightarrow K}}{(s|np)|np}}{
\infer{\overset{c^{(B\rightarrow 2)\rightarrow (1\rightarrow A)\rightarrow B\rightarrow A}}{(s|np)|np}}{\textit{caught}} &
\infer{\overset{(\textit{and}\, a)^{((E\rightarrow E)\rightarrow(F\rightarrow F)\rightarrow
2\rightarrow I)\rightarrow(L\rightarrow
4)\rightarrow(I\rightarrow K)\rightarrow L\rightarrow K}}{((s|np)|np)\,|\,((s|np)|np)}}{\infer{\overset{\textit{and}^{((G\rightarrow G)\rightarrow(H\rightarrow H)\rightarrow J\rightarrow 3)\rightarrow((E\rightarrow E)\rightarrow(F\rightarrow F)\rightarrow
2\rightarrow I)\rightarrow(L\rightarrow
J)\rightarrow(I\rightarrow K)\rightarrow L\rightarrow K}}{(((s|np)|np)\,|\,((s|np)|np))\,|\,((s|np)|np)}}{\textit{and}}
& \infer{\overset{a^{(D\rightarrow 4)\rightarrow(3\rightarrow C)\rightarrow D\rightarrow C}}{(s|np)|np}}{\textit{ate}}}}
& \infer{\overset{f^{5\rightarrow 4}}{np}}{\textit{a fish}}}}
$$
\end{sideways}
\end{center}
\caption{Proof of ``John caught and ate a fish'' (simplified)}
\label{fig:fish}
\end{figure}
As we can see from the proof in Figure~\ref{fig:fish}, this lexical type allows us to derive
``John caught and ate a fish'' with the correct semantics. The proof
has been slightly simplified by using distinct variables for the words
instead of complex lambda terms (ie.\ we have not done lexical
substitution). This has the advantage that we can use the resulting
lambda term for computing the semantics as well, for which we use the
following (standard) semantic substitutions\footnote{To keep this example simple, we have
treated ``a fish" as an individual constant instead of a quantified
noun phrase, since quantification is irrelevant for this
example.}. We can obtain the prosodic lambda terms from the
principal types and the string positions (eg.\ $(2\rightarrow 1) \vdash (B\rightarrow 2)
\rightarrow (1\rightarrow A) \rightarrow B\rightarrow A$ for ``caught'', which is the standard
transitive verb principal type we have seen before).
The semantic terms below are all standard.
\begin{align*}
\textit{and} &= \lambda \textit{TV1} \lambda \textit{TV2}\lambda
y \lambda x. ((\textit {TV1}\, y)\, x) \wedge ((\textit{TV2}\, y)\, x) \\
j &= \textit{john'} \\
f &= \textit{a\_fish'} \\
c &= \textit{caught'} \\
a &= \textit{ate'}
\end{align*}
Unfortunately, this analysis of ``and'' also make the
(rather catastrophic) prediction that ``John caught and ate a fish'' has a
second reading which can be paraphrased as ``John caught a fish and a
fish ate John''. This reading is easy to miss when we look only at
eta-short proofs, since the key point of this second derivation
involves switching the two arguments of the transitive verb, as
shown in Figure~\ref{fig:fishb}.\footnote{The term $\lambda f. \lambda x.\lambda y. ((f\, y)\,
x)$ which switches subject and object is of course the \textbf{C} combinator we have already seen in
Section~\ref{sec:exc}. It commutes the two
arguments of a function $f$, and the proof shown in
Figure~\ref{fig:fishb} has a subproof which computes
$\textbf{C}a\equiv \lambda x.\lambda y. (a\, y)\,x$.} The crux
of this second proof is that swapping the two arguments of ``ate'' is
a purely local operation which has no visible effects on the word
order: the only difference between the proof in Figure~\ref{fig:fish}
and the proof in
Figure~\ref{fig:fishb} is in the subproof with undischarged hypothesis
``ate'' (with term $a$ resp.\ $\lambda x. \lambda y. (a\, y)\,x$).
\editout{
$$
\infer[\multimap I_1]{\lambda x. \lambda y. ((a\, y)\, x) :np \multimap np \multimap s}{\infer[\multimap
I_2]{\lambda y. ((a\, y)\, x): np\multimap s}{\infer[\multimap E]{((a\, y)\, x):s}{[x:np]^1 & \infer[\multimap
E]{(a\, y):np \multimap s}{[y:np]^2 & a:np \multimap np \multimap s }}}}
$$
eta-expanded
$$
((\textit{and}\,
\lambda y. \lambda x. ((a\, y)\, x)\, c)\, f)\, j
$$
}
\editout{
$$
\infer{s}{\infer{np}{\textit{John}} &
\infer{np\rightarrow s}{
\infer{np\rightarrow np\rightarrow s}{
\infer{np\rightarrow np\rightarrow s}{\textit{caught}} &
\infer{a2}{\infer{a1}{\textit{and}} & \infer[I_2]{np\rightarrow
np\rightarrow s}{\infer[I_1]{np\rightarrow s}{\infer{s}{\infer{[np]^1}{x} &\infer{np\rightarrow
s}{\infer{np\rightarrow
np\rightarrow s}{\textit{ate}} & \infer{[np]^2}{y}}}}}}}
& \infer{np}{\textit{a fish}}}}
$$
$$
\infer{s}{\infer{np}{\textit{John}} &
\infer{np\rightarrow s}{
\infer{np\rightarrow np\rightarrow s}{
\infer{np\rightarrow np\rightarrow s}{\textit{caught}} &
\infer{a2}{\infer{a1}{\textit{and}} & \infer[I_1]{np\rightarrow
np\rightarrow s}{\infer[I_2]{np\rightarrow s}{\infer{s}{\infer{[np]^1}{x} &\infer{np\rightarrow
s}{\infer{np\rightarrow
np\rightarrow s}{\textit{ate}} & \infer{[np]^2}{y}}}}}}}
& \infer{np}{\textit{a fish}}}}
$$
}
\begin{figure}
\begin{center}
\begin{sideways}
$$
\infer{\overset{((((\textit{and}\,
a)\, c)\, f)\, j)^{5\rightarrow 0}}{s}}{\infer{np^{1\rightarrow 0}}{\textit{John}} &
\infer{\overset{(((\textit{and}\,
a)\, c)\, f)^{(1\rightarrow
K)\rightarrow 5\rightarrow K}}{s|np}}{
\infer{\overset{((\textit{and}\,
a)\, c)^{(L\rightarrow 4)\rightarrow (1\rightarrow
K)\rightarrow L\rightarrow K}}{(s|np)|np}}{
\infer{\overset{c^{(B\rightarrow 2)\rightarrow (1\rightarrow A)\rightarrow B\rightarrow A}}{(s|np)|np}}{\textit{caught}} &
\infer{\overset{(\textit{and}\, a)^{((E\rightarrow E)\rightarrow(F\rightarrow F)\rightarrow
2\rightarrow I)\rightarrow(L\rightarrow
4)\rightarrow(I\rightarrow K)\rightarrow L\rightarrow K}}{((s|np)|np)\,|\,((s|np)|np)}}{\infer{\overset{\textit{and}^{((G\rightarrow G)\rightarrow(H\rightarrow H)\rightarrow J\rightarrow 3)\rightarrow((E\rightarrow E)\rightarrow(F\rightarrow F)\rightarrow
2\rightarrow I)\rightarrow(L\rightarrow
J)\rightarrow(I\rightarrow K)\rightarrow L\rightarrow K}}{(((s|np)|np)\,|\,((s|np)|np))\,|\,((s|np)|np)}}{\textit{and}}
& \infer{\overset{(\lambda x. \lambda y. ((a\,y)\,x))^{(3\rightarrow C)\rightarrow(D\rightarrow
4)\rightarrow D\rightarrow C}}{(s|np)|np}}{\infer{\overset{(\lambda y. ((a\,y)\,x))^{(D\rightarrow
4)\rightarrow D\rightarrow C}}{s|np}}{\infer{\overset{((a\,y)\,x)^{D\rightarrow C}}{s}}{\infer{\overset{(a\,y)^{(3\rightarrow C)\rightarrow D\rightarrow C}}{s|np}}{\infer{\overset{a^{(D\rightarrow 4)\rightarrow(3\rightarrow C)\rightarrow D\rightarrow
C}}{(s|np)|np}}{\textit{ate}} &
\overset{y^{D\rightarrow 4}}{np} } & \overset{x^{3\rightarrow C}}{np}}}}
}}
& \infer{\overset{f^{5\rightarrow 4}}{np}}{\textit{a fish}}}}
$$
\end{sideways}
\end{center}
\caption{Proof of ``John caught and ate a fish'' with semantics ``John
caught a fish and a fish ate John'' (simplified).}
\label{fig:fishb}
\end{figure}
As shown in the figure, the second proof computes the following ``deep structure''.
$$(((\textit{and}\,
(\textbf{C} a))\, c)\, f)\, j = (((\textit{and}\,
\lambda x. \lambda y. ((a\, y)\, x)\, c)\, f)\, j$$
In a similar way, we can obtain a third and a fourth reading,
corresponding the string ``John caught and ate a fish'' but to the meanings ``A fish caught John and John ate a
fish'' and ``A fish caught and ate John'' respectively, as follows.
\begin{align*}
(((\textit{and}\,
a)\, (\textbf{C} c))\, f)\, j &= (((\textit{and}\, a)\, \lambda v.\lambda w. (c\, w)\,
v)\, f)\, j\\
(((\textit{and}\,
(\textbf{C} a))\, (\textbf{C} c))\, f)\, j &= (((\textit{and}\,
\lambda x. \lambda y. ((a\, y)\, x)\, \lambda v.\lambda w. (c\, w)\,
v)\, f)\, j
\end{align*}
The problem is that though we would \emph{want} the two $(s|np)|np$
arguments of ``and'' to be transitive verbs, they \emph{mean} ``a
sentence missing two $\textit{np}$ arguments anywhere'', which is what
causes the problems with commutativity.
We can again remedy this by adding new lexical entries, for example
choosing $tv$ for the two transitive verbs and $(tv\backslash ((np\backslash
s)/np))/tv$ for the conjunction, but this would mean adding several
other lexical entries to analyse sentences like ``John has understood and
will probably implement Dijkstra's algorithm'', which are handled by
the Lambek calculus analysis --- since ``has understood'' and ``will
probably implement'' can both be analysed as $(np\backslash s)/np$ ---
but not by the new atomic $tv$ analysis. So adding lexical entries is
not only inelegant and an admittance of descriptive inadequacy, but
such additions can cascade throughout the grammar.
I would seem that another simple potential solution would be to add
case to lambda grammars. While adding case to
first-order linear logic is something we can do essentially for free
using extra arguments, adding case to lambda grammars at least
complicates either the grammars or the types. In addition, though case
would exclude the subject-object
swaps we have seen in this section, is is easy to see this would not be a real
solution, because the sentences in \ref{subjex} below are all sentences
missing a subject/nominative $np$, those in \ref{objex} sentences
missing an object/accusative $np$
and those in \ref{tvex} sentences missing both a subject and an
object (for clarity, the missing subjects and objects have been shown
as $[]_s$ and $[]_o$ respectively). So while adding case excludes some bad derivations, we would still
predict sentences like ``*Sue likes Mary and John saw the man whom
likes'' is grammatical (with meaning ``Sue likes Mary and John saw the man whom Sue
likes.''), that `` *John saw the friend of who lives in Paris and
Ted likes Sue'' is grammatical and means ``John saw the friend of Sue who lives in Paris
and Ted likes Sue'' and that ``*Sue John believes avoids but Ted saw
whom kissed Peter'' is grammatical and means ``John believes Sue avoids
Peter but Ted saw Peter whom Sue kissed''.
\ex.\label{subjex}
\a. $[]_{s}$ likes Mary.
\b. John believes $[]_s$ left.
\b. John saw the man whom $[]_s$ likes.
\ex.\label{objex} \a. John likes $[]_o$.
\b. John saw the friend of $[]_o$ who lives in Paris.
\b. Ted gave $[]_o$ flowers.
\ex.\label{tvex} \a. $[]_s$ gave $[]_o$ flowers.
\b. John believes $[]_s$ avoids $[]_o$.
\b. John saw $[]_o$ whom $[]_s$ kissed.
While it would certainly be possible to appeal to island constraints
or other independently motivated mechanisms to exclude coordination of
the phrases listed above, it seems that use case for this purpose is inherently on
the wrong track: it uses lambda terms to encode word order for
negative implications and a cascade of stop-gap solutions to
constrain word order for positive implications. As the examples above
make clear, for coordination, we
don't coordinate (partial) constituents which have the same case
marking, but rather those which have the same \emph{structure} and
Lambek calculus formulas, such as $(np\backslash s)/np$ for transitive verb conjunction, are a good proxy for this notion of the same structure.
Though we have given an in-depth analysis only of transitive verb
conjunction, other conjunctions of complex types (adjectives,
intransitive verbs, etc.) suffer similar problems. In all cases, the
lambda grammar analysis is between a rock and a hard place, suffering
either from overgeneration (as the adverb case) or from bizarre
readings (as in the transitive verb conjunction case).
\subsubsection*{Gapping}
As a last problem case, the standard (multimodal) categorial grammar analysis of gapping
\cite{cgellipsis}, of which we have seen the hybrid version in
Section~\ref{sec:string}, does not fare any better when we try to
translate it into lambda grammar. The analysis of a sentence like
\ex. John studies logic and Charles phonetics.
\noindent would assign ``and'' the formula.\editout{
$$((np\multimap np \multimap s)
\multimap s) \multimap ((np\multimap np \multimap s)
\multimap s) \multimap (np\multimap np \multimap s)
\multimap s$$
}
$$
((s|((s|np)|np))\,|\,(s|((s|np)|np)))\,|\,(s|((s|np)|np))
$$
The idea behind the analysis of \citeasnoun{cgellipsis} is that
``and'' takes first two sentences missing a transitive verb as its
arguments, then a transitive verb to produce a sentence by placing the
transitive verb back to its normal place in the first argument (which
is the sentences to its left missing a transitive verb) and using the
empty string instead of the transitive verb in the second sentence. In
short, it uses a quantifying-in analysis for the transitive verb in
the sentence to the left and an extraction analysis for the
``missing'' transitive in the sentence to the right. What is nice
about this analysis, is that we use a normal coordination formula
for ``and'', an instance of the schema $(X|X)|X$ with (in this case)
$X = (s|np)|np$. Since the analysis of \citeasnoun{cgellipsis} uses a combination of quantifying in and
extraction, it is tempting to think that the lambda grammar analysis
is unproblematic. However, they key point of the analysis is that we
need both extraction and in situ binding for a complex formula, a
transitive verb, though unlike for the coordination case it occurs in
a negative position in the gapping coordination type. Let's
investigate the possibilities.
The formula for transitive verb gapping produces the (simplified and reduced) inhabitation machine
shown in Figure~\ref{fig:inhabg}.
\begin{figure}
\begin{center}
\begin{tikzpicture}[node distance=1.5cm]
\node[draw] (SIGMA) {$\ \sigma\ $};
\path[->] (SIGMA) edge [loop above] node[above] {$\textit{and}$} ();
\node (Z) [below of=SIGMA] {$z$};
\node (STV1) [left of=Z] {$\textit{STV1}$};
\node (dmy1) [below of=STV1] {};
\node[draw] (SSS) [left of=dmy1] {$(\sigma\rightarrow\sigma)\rightarrow(\sigma\rightarrow\sigma)\rightarrow\sigma\rightarrow\sigma$};
\node[draw] (ZZZ) [below of=SSS] {$\ \sigma\ $};
\node (dmy3) [below of=ZZZ] {};
\node (XXX) [left of=dmy3] {$x$};
\node (TV1) [right of=dmy3] {$\textit{TV}$};
\node (TV2) [right of=Z] {$\textit{STV2}$};
\path[->] (SSS) edge node[left] {$\lambda \textit{O1} \lambda \textit{S1} \lambda x$} (ZZZ);
\draw[->] (SIGMA.west) -- ([xshift=-3pt] STV1.north);
\draw[->] ([xshift=3pt] STV1.north) -- (SIGMA.south west);
\draw[->] (SIGMA.east) -- ([xshift=3pt] TV2.north);
\draw[->] ([xshift=-3pt] TV2.north) -- (SIGMA.south east);
\draw[->] (SIGMA.south) -- (Z.north);
\node (dmy4) [below of=TV1] {};
\node[draw] (OBJ1) [right of=dmy4] {$\sigma\rightarrow\sigma$};
\node[draw] (SUBJ1) [left of=dmy4] {$\sigma\rightarrow\sigma$};
\node[draw] (SUBJ11) [below of=SUBJ1] {$\sigma$};
\node[draw] (OBJ11) [below of=OBJ1] {$\sigma$};
\node (X) [below of=SUBJ11] {$v$};
\node (Y) [below of=OBJ11] {$w$};
\draw[->] (ZZZ.east) -- ([xshift=3pt] TV1.north);
\draw[->] ([xshift=-3pt] TV1.north) -- (ZZZ.south east);
\path[->] (SUBJ11) edge [loop left] node[left] {$\textit{S1}$} ();
\path[->] (OBJ11) edge [loop right] node[right] {$\textit{O1}$} ();
\path[->] (SUBJ1) edge node[left] {$\lambda v$} (SUBJ11);
\path[->] (OBJ1) edge node[left] {$\lambda w$} (OBJ11);
\draw[->] (STV1) -- (SSS);
\draw[->] (ZZZ) --(XXX);
\draw[->] (TV1) -- (OBJ1);
\draw[->] (TV1) -- (SUBJ1);
\draw[->] (OBJ11) -- (Y);
\draw[->] (SUBJ11) -- (X);
\node (dmy2) [below of=TV2] {};
\node[draw] (SUBJ2) [right of=dmy2] {$(\sigma\rightarrow\sigma)\rightarrow(\sigma\rightarrow\sigma)\rightarrow\sigma\rightarrow\sigma$};
\node[draw] (SUBJ22) [below of=SUBJ2] {$\sigma$};
\node (V) [below of=SUBJ22] {$y$};
\draw[->] (TV2) -- (SUBJ2);
\draw[->] (SUBJ22) -- (V);
\path[->] (SUBJ2) edge node[left] {$\lambda \textit{O2} \lambda \textit{S2} \lambda y$} (SUBJ22);
\path[->] (SUBJ22) edge [loop left] node[left] {$\textit{S2}$} ();
\path[->] (SUBJ22) edge [loop right] node[right] {$\textit{O2}$} ();
\end{tikzpicture}
\end{center}
\caption{Simplified inhabitation machine for gapping.}
\label{fig:inhabg}
\end{figure}
As before, the prefix $\lambda \textit{STV2}. \lambda \textit{STV1}.
\lambda \textit{TV}. \lambda z.$ has been remove from the figure;
$\textit{STV1}$ denotes the sentence missing a transitive verb to the
left of ``and'' (``John logic'' in our case) and $\textit{STV2}$
denotes the sentence missing a transitive verb to the right of ``and''
(``Charles phonetics'' in our case), $\textit{TV}$ the transitive verb
(here: ``studies'') and $z$ the end of the string.
We can obtain the full combinatorics by identifying all nodes with the
same type. The current reduced graph emphasizes the reasonable
lambda terms: for example, $\textit{TV}$ can only be an argument of
$\textit{STV1}$, corresponding to the the quantifying-in analysis
producing the desired word order ``John
studies logic'', similarly, the first argument of the transitive verb
has been restricted to the subject and the second argument to the object.
In fact, getting the word order and semantics right leaves a unique
lambda term --- this is just the term from \cite{bourreau13ellipse},
where the subterm $(\textit{STV1}\, \textit{TV})$ has been eta-expanded.\footnote{The eta-short term looks as follows.
\begin{align*}
\lambda \textit{STV2}. \lambda \textit{STV1}.
\lambda \textit{TV}. \lambda z. (&(\textit{STV1}\,\textit{TV})\\
& (\textit{and}\ (\textit{STV2}\, \lambda \textit{O2} \lambda \textit{S2} \lambda
y. \textit{S2}\, (\textit{O2}\, y))\, z))
\end{align*}}
\begin{align*}
\lambda \textit{STV2}. \lambda \textit{STV1}.
\lambda \textit{TV}. \lambda z. (&(\textit{STV1}\,\lambda \textit{O1}
\lambda \textit{S1} \lambda x.(((\textit{TV}\, \lambda w.\textit{O1}\,
w)\, \lambda v.\textit{S1}\, v))\, x)\\
& (\textit{and}\ (\textit{STV2}\, \lambda \textit{O2} \lambda \textit{S2} \lambda
y. \textit{S2}\, (\textit{O2}\, y))\, z))
\end{align*}
The principal type of this term is.
\begin{align*}
(((L\rightarrow K) \rightarrow (K \rightarrow M) &\rightarrow L \rightarrow M) \rightarrow J \rightarrow 4) \rightarrow \\
(((D\rightarrow C) \rightarrow (F\rightarrow E) &\rightarrow H \rightarrow G) \rightarrow 3 \rightarrow I) \rightarrow \\
((D \rightarrow C) \rightarrow (F \rightarrow E) &\rightarrow H \rightarrow G) \rightarrow J \rightarrow I
\end{align*}
Given this lambda term and principal type, the we can derive the
correct word order and semantics as shown in Figure~\ref{fig:jsl}.
We have again abbreviated the proof, using the following abbreviations
for readability.
\begin{align*}
X &= s|((s|np)|np) \\
\alpha &= ((L\rightarrow K)\rightarrow (K\rightarrow M) \rightarrow L\rightarrow M)\rightarrow J\rightarrow 4 \\
\beta &= (D\rightarrow C)\rightarrow (F\rightarrow E)\rightarrow H\rightarrow G\\
\end{align*}
We have also performed the substitutions necessary for the $\rightarrow E$
rules directly on the hypotheses of the proof. We obtain the semantics by
substituting the following terms for the constants in the lambda term
computed for this proof.
\begin{align*}
\textit{and} &= \lambda \textit{STV1} \lambda \textit{STV2} \lambda
\textit{TV}. (\textit {STV1} \textit{TV}) \wedge (\textit{STV2}
\textit{TV}) \\
j &= \textit{john'} \\
l &= \textit{logic'} \\
c &= \textit{charles'} \\
p &= \textit{phonetics'} \\
s & = \textit{studies'}
\end{align*}
But again, there is an alternative
proof, shown in Figure~\ref{fig:jslb}. This proof swaps both the arguments of $P$ and the
two abstractions of $s$ (``studies''). The net result is that the left
conjunct stays as before, syntactically and semantically, since the
two swaps cancel out against each other. Now the
right conjunct has its arguments swapped in the semantics only, giving the absurd reading ``John
studies logic and phonetics studies Charles''.
\newcommand{\bigforma}{\overset{\textit{and}^{\alpha\rightarrow (\beta\rightarrow 3\rightarrow
I)\rightarrow \beta\rightarrow J\rightarrow I}}{(X|X)|X}}
\newcommand{\bigformb}{\overset{(\textit{and}\, (\lambda
Q. (Q\,p)\,c))^{(\beta \rightarrow 3\rightarrow I) \rightarrow \beta \rightarrow 6\rightarrow I}}{X|X}}
\editout{
\begin{align*}
b1 &= ((np \rightarrow np\rightarrow s)\rightarrow s) \rightarrow ((np \rightarrow np\rightarrow s)\rightarrow s) \rightarrow (np \rightarrow
np \rightarrow s) \rightarrow s \\
b2 &= ((np \rightarrow np\rightarrow s)\rightarrow s) \rightarrow (np \rightarrow
np \rightarrow s) \rightarrow s\\
\end{align*}
$$
\infer{s}{
\infer{(np \rightarrow np\rightarrow s)\rightarrow s}{
\infer[I_1]{(np \rightarrow np \rightarrow s)\rightarrow s}{\infer{s}{\infer{np}{\textit{John}} &
\infer{np\rightarrow s}{\infer{[np \rightarrow np\rightarrow s]^1}{P} &
\infer{np}{\textit{logic}}}}} &
\infer{\bigformb}{
\infer{\bigforma}{\textit{and}} &
\infer[I_2]{(np \rightarrow np \rightarrow s)\rightarrow s}{\infer{s}{\infer{np}{\textit{Charles}} &
\infer{np\rightarrow s}{\infer{[np \rightarrow np\rightarrow s]^2}{Q} &
\infer{np}{\textit{phonetics}}}}}}} &
\infer{np\rightarrow np\rightarrow s}{\textit{studies}}}
$$
$$
\infer{s}{
\infer{(np \rightarrow np\rightarrow s)\rightarrow s}{
\infer[I_1]{(np \rightarrow np \rightarrow s)\rightarrow s}{\infer{s}{\infer{np}{\textit{John}} &
\infer{np\rightarrow s}{\infer{[np \rightarrow np\rightarrow s]^1}{P} &
\infer{np}{\textit{logic}}}}} &
\infer{\bigformb}{
\infer{\bigforma}{\textit{and}} &
\infer[I_2]{(np \rightarrow np \rightarrow s)\rightarrow s}{\infer{s}{\infer{np}{\textit{Charles}} &
\infer{np\rightarrow s}{\infer{[np \rightarrow np\rightarrow s]^2}{Q} &
\infer{np}{\textit{phonetics}}}}}}} &
\infer{np\rightarrow np\rightarrow s}{\textit{studies}}}
$$
}
\begin{figure}
\begin{center}
\begin{sideways}
$$
\infer{\overset{(((\textit{and}\, (\lambda
Q. (Q\,p)\,c))\, (\lambda P. (P\,l)\,j))\, (\lambda yx. (s\,y)\,x))^{6\rightarrow 0}}{\vphantom{|}s}}{
\infer{\overset{((\textit{and}\, (\lambda
Q. (Q\,p)\,c))\, (\lambda P. (P\,l)\,j))^{((3\rightarrow 2)\rightarrow(1\rightarrow 0) \rightarrow
3\rightarrow I)\rightarrow 6\rightarrow I}}{s|((s|np)|np)}}{
\infer{\overset{(\lambda P. (P\,l)\,j)^{((3\rightarrow 2)\rightarrow(1\rightarrow 0)\rightarrow
B'\rightarrow A')\rightarrow B'\rightarrow
A'}}{s|((s|np)|np)}}{\!\!\!\!\!\!\infer{\overset{((P\,l)\,j)^{B'\rightarrow
A'}}{\vphantom{|}s}\!\!\!\!\!\!}{\infer{\overset{j^{\vphantom{(P)^{1\rightarrow
0}}1\rightarrow 0}}{\vphantom{|}np}}{\textit{John}} &
\infer{\overset{(P\,l)^{(1\rightarrow 0)\rightarrow B'\rightarrow
A'}}{s|np}}{\overset{P^{(3\rightarrow 2)\rightarrow (1\rightarrow 0)\rightarrow B'\rightarrow A'}}{(s|np)|np} &
\infer{\overset{l^{3\rightarrow 2}}{\vphantom{|}np}}{\textit{logic}}}}} &
\infer{\bigformb}{
\infer{\bigforma}{\textit{and}} &
\infer{\overset{(\lambda Q. (Q\,p)\,c)^{((6\rightarrow 5)\rightarrow(5\rightarrow 4)\rightarrow
D'\rightarrow C')\rightarrow D'\rightarrow C'}}{s|((s|np)|np)}}{\infer{\overset{((Q\,p)\,c)^{D'\rightarrow C'}}{\vphantom{|}s}}{\infer{\overset{c^{5\rightarrow 4}}{\vphantom{|}np}}{\textit{Charles}} &
\infer{\overset{(Q\, p)^{(5\rightarrow 4)\rightarrow D'\rightarrow C'}}{s|np}}{\overset{Q^{(6\rightarrow 5)\rightarrow (5\rightarrow 4)\rightarrow D'\rightarrow C'}}{(s|np)|np} &
\infer{\overset{p^{6\rightarrow 5}}{np}}{\textit{phonetics}}}}}}} &
\infer{\overset{(\lambda yx. (s\,y)\,x)^{(B\rightarrow
2)\rightarrow(1\rightarrow A)\rightarrow B\rightarrow A}}{(s|np)|np}}{\infer{\overset{(\lambda
x.(s\,y)\,x)^{(1\rightarrow A)\rightarrow B\rightarrow A}}{s|np}}{\infer{\overset{((s\,y)\, x)^{B\rightarrow A}}{\vphantom{|}s}}{\overset{x^{1\rightarrow A}}{\vphantom{|}np} & \infer{\overset{(s\, y)^{(1\rightarrow
A)\rightarrow B\rightarrow A}}{s|np}}{\infer{\overset{s^{(B\rightarrow 2)\rightarrow (1\rightarrow
A)\rightarrow B\rightarrow A}}{(s|np)|np}}{\textit{studies}} &
\overset{y^{B\rightarrow 2}}{\vphantom{|}np} }}}}}
$$
\end{sideways}
\end{center}
\caption{Proof of "John studies logic and Charles phonetics".}
\label{fig:jsl}
\end{figure}
\begin{figure}
\begin{center}
\begin{sideways}
$$
\infer{\overset{(((\textit{and}\, (\lambda
Q. (Q\,p)\,c))\, (\lambda P. (P\,j)\,l))\, (\lambda xy. (s\,y)\,x))^{6\rightarrow 0}}{s}}{
\infer{\overset{((\textit{and}\, (\lambda
Q. (Q\,p)\,c))\, (\lambda P. (P\,j)\,l))^{((1\rightarrow 0)\rightarrow(3\rightarrow 2) \rightarrow
3\rightarrow I)\rightarrow 6\rightarrow I}}{s|((s|np)|np)}}{
\infer{\overset{(\lambda P. (P\,j)\,l)^{((1\rightarrow 0)\rightarrow(3\rightarrow 2)\rightarrow
B'\rightarrow A')\rightarrow B'\rightarrow
A'}}{s|((s|np)|np)}}{
\infer{\overset{((P\,j)\,l)^{B'\rightarrow A'}}{s}}{
\infer{\overset{(P\,j)^{(3\rightarrow 2)\rightarrow B'\rightarrow A'}}{s|np}}{
\infer{\overset{j^{1\rightarrow 0}}{np}}{\textit{John}} &
\overset{P^{(1\rightarrow 0)\rightarrow (3\rightarrow 2)\rightarrow B'\rightarrow A'}}{(s|np)|np}}
& \infer{\overset{l^{3\rightarrow 2}}{np}}{\textit{logic}}
}
} &
\infer{\bigformb}{
\infer{\bigforma}{\textit{and}} &
\infer{\overset{(\lambda Q. (Q\,p)\,c)^{((6\rightarrow 5)\rightarrow(5\rightarrow 4)\rightarrow
D'\rightarrow C')\rightarrow D'\rightarrow C'}}{s|((s|np)|np)}}{\infer{\overset{((Q\,p)\,c)^{D'\rightarrow C'}}{s}}{\infer{\overset{c^{5\rightarrow 4}}{np}}{\textit{Charles}} &
\infer{\overset{(Q\, p)^{(5\rightarrow 4)\rightarrow D'\rightarrow C'}}{s|np}}{\overset{Q^{(6\rightarrow 5)\rightarrow (5\rightarrow 4)\rightarrow D'\rightarrow C'}}{(s|np)|np} &
\infer{\overset{p^{6\rightarrow 5}}{np}}{\textit{phonetics}}}}}}} &
\infer{\overset{(\lambda xy. (s\,y)\,x)^{(1\rightarrow
A)\rightarrow(B\rightarrow 2)\rightarrow B\rightarrow A}}{(s|np)|np}}{\infer{\overset{(\lambda
y.(s\,y)\,x)^{(B\rightarrow 2)\rightarrow B\rightarrow A}}{s|np}}{\infer{\overset{((s\,y)\, x)^{B\rightarrow A}}{s}}{\overset{x^{1\rightarrow A}}{np} & \infer{\overset{(s\, y)^{(1\rightarrow
A)\rightarrow B\rightarrow A}}{s|np}}{\infer{\overset{s^{(B\rightarrow 2)\rightarrow (1\rightarrow
A)\rightarrow B\rightarrow A}}{(s|np)|np}}{\textit{studies}} &
\overset{y^{B\rightarrow 2}}{np} }}}}}
$$
\end{sideways}
\end{center}
\caption{Proof of "John studies logic and Charles phonetics" with
semantics ``John studies logic and phonetics studies Charles''.}
\label{fig:jslb}
\end{figure}
\editout{
$$
\infer{s}{
\infer{(np \rightarrow np\rightarrow s)\rightarrow s}{
\infer[I_1]{(np \rightarrow np \rightarrow s)\rightarrow s}{\infer{s}{\infer{np}{\textit{logic}} &
\infer{np\rightarrow s}{\infer{[np \rightarrow np\rightarrow s]^1}{P} &
\infer{np}{\textit{John}}}}} &
\infer{\bigformb}{
\infer{\bigforma}{\textit{and}} &
\infer[I_2]{(np \rightarrow np \rightarrow s)\rightarrow s}{\infer{s}{\infer{np}{\textit{Charles}} &
\infer{np\rightarrow s}{\infer{[np \rightarrow np\rightarrow s]^2}{Q} &
\infer{np}{\textit{phonetics}}}}}}} &
\infer[I_3]{np\rightarrow np\rightarrow s}{\infer[I_4]{np\rightarrow s}{\infer{s}{\infer{[np]^3}{x} & \infer{np\rightarrow
s}{\infer{np\rightarrow np\rightarrow s}{\textit{studies}} & \infer{[np]^4}{y}}}}}}
$$
}
Since this article is already rather long, we cannot treat the other problem cases mentioned at the start of
Section~\ref{inadeq}. However, the examples which have been treated in
detail serve as a blueprint to constructing similar problems for the
additional listed problem cases.
In all cases, the fundamental asymmetry of
lambda grammars means we have insufficient tools at our disposal to constrain the word order for positive
implications and that the best possible approximations are inadequate both
syntactically and semantically.
\subsection{Solutions for lambda grammars}
\label{solutions}
Given the descriptive challenges for lambda grammars, it seems natural to ask how
lambda grammars could evolve to rise to these challenges.
\begin{enumerate}
\item \emph{Stasis.} Keeping the formalism and the analyses as
they are is a possible, if not a very attractive solution, since it
would require us to significantly tone down the ambitions of the
syntax-semantics interface of the formalism, thereby losing one of
the attractive aspects of categorial grammars.
We can also choose to embrace descriptive inadequacy and use the
result from \cite{busz96}, which, as discussed in Section~\ref{sec:mainthm},
translates
Lambek grammars into AB grammars while preserving the semantics, though at the price of an
explosion in lexicon size (as Pentus', 1995 original proof) --- to obtain
at least the most of the
coverage of hybrid type-logical grammars directly within
lambda grammars (though this presupposes strict separation, as required
for the application of Lemma~\ref{lem:hybridlambda}). This would save lambda grammars empirically,
incorporating the syntax-semantics interface of the hybrid system, but we
would then have a combinatorial explosion followed by an
NP-complete
problem (according to Lemma~\ref{lem:nplambda}). Given that the original Pentus proof, with $O(|G | n^3)$ complexity for
some colossal $| G |$, never resulted in fast, practical parsers for
the Lambek calculus because of the grammar size constant, having
a similar constant for an NP-complete problem does not bode well for
parsing the resulting grammar.
So it seems we have two unappealing options here: give up --- or
significantly reduce the ambition of --- the syntax-semantics
interface or give up actually parsing lambda grammars.
\item Change the \emph{terms} and/or their
\emph{interpretation}. The lambda grammars discussed above produce
strings. Several authors have looked at lambda grammars which
generate different types of structures, such as trees \cite{muskens01lfg,groote02tag}. When we
generate trees, we can add a separate yield algebra which tells us
how to interpret the possible word orders generated by a given
tree. This ``multiple transduction'' approach has several other
instances and though it is conceivable that such multiple transductions may help alleviate some of the \emph{symptoms},
it does not address their root cause, which is the asymmetry of the
system.
What we need is a system which can reject ``candidate derivations''
which have been computed at the previous level.
Such a solution can be found in \cite{muskens97comb}, who produces first-order logic
formulas and adds a separate theorem-prover component (which is
essentially a model-builder for multimodal categorial
grammars). Muskens' solution would solve the problems with
lambda grammars by essentially generating potential derivations and
asking a multimodal grammar if these derivations are valid. However,
this setup does not seem to have any benefits over a direct
multimodal implementation and suffers from the same complexity
problems as multimodal categorial grammars. \citeasnoun{pp10acg},
discussed below since they change the types in addition to the terms, also fall into this
category.
Another potential solution in this family would be to add
\emph{term equations} (and corresponding reductions) to the lambda
calculus. However, it is unclear what sort of form such a solution
would take.
\editout{
A simple solution to reject some undesired derivations would be to add
case to lambda grammars. While this would include the subject-object
swaps which were the cause of catastrophic overgeneration for
transitive verb conjunction and gapping, is is easy to see this would not be a real
solution, because the sentences in \ref{subjex} below are all sentences
missing a subject/nominative $np$, those in \ref{objex} sentences
missing an object/accusative $np$
and those in \ref{tvex} sentences missing both a subject and an
object (for clarity, the missing subjects and objects have been shown
as $[]_s$ and $[]_o$ respectively). So while adding case excludes some bad derivations, we would still
predict sentences like ``*Sue likes Mary and John saw the man whom
likes'' is grammatical (with meaning ``Sue likes Mary and John saw the man whom Sue
likes.''), that `` *John saw the friend of who lives in Paris and
Ted likes Sue'' is grammatical and means ``John saw the friend of Sue who lives in Paris
and Ted likes Sue'' and that ``*Sue John believes avoids but Ted saw
whom kissed Peter'' is grammatical and means ``John believes Sue avoids
Peter but Ted saw Peter whom Sue kissed''.
\ex.\label{subjex}
\a. $[]_{s}$ likes Mary.
\b. John believes $[]_s$ left.
\b. John saw the man whom $[]_s$ likes.
\ex.\label{objex} \a. John likes $[]_o$.
\b. John saw the friend of $[]_o$ who lives in Paris.
\b. Ted gave $[]_o$ flowers.
\ex.\label{tvex} \a. $[]_s$ gave $[]_o$ flowers.
\b. John believes $[]_s$ avoids $[]_o$.
\b. John saw $[]_o$ whom $[]_s$ kissed.
While it would certainly be possible to appeal to island constraints
or other independently motivated mechanisms to exclude coordination of
the phrases listed above, it seems that this approach is inherently on
the wrong track: it uses lambda terms to encode word order for
negative implications and a cascade of stop-gap solutions to
constrain word order for positive implications.}
As we have seen in the treatment of transitive verb conjunctions,
adding case offers (at best) a partial solution by attacking the
symptoms rather than the underlying cause of the problem. It should also be noted that
the only solution of this kind which has been worked out in any detail
uses dependent types and this complicates the \emph{types} as well as the terms
\cite{pp10acg,pompigne13}, which moves us to the next point.
\item Change the \emph{types}. Various authors \cite{maarek,pp10acg} have looked at extending the type theory of
lambda grammars beyond the simply typed lambda calculus. Of these
extensions, dependent types seem well-suited to the challenges posed
in this paper, though they would need to be added on a much larger
scale than previously assumed and they would complicate the
type/term calculus and its mathematical properties considerably.
An alternative solution, proposed in the context of linear grammars \cite{worth14coord}
uses subtyping combined with restrictions on the form
of subterm cooccurrences. It is unclear to me at the moment whether this
type of treatment is equivalent to other proposals (eg.\ those of
hybrid type-logical grammars) and whether it corresponds to a natural
fragment of first-order linear logic.
\end{enumerate}
Is seems that the easiest way to fix the inadequacies of
lambda grammars would by to extend the system to hybrid type-logical
grammar: existing linguistic analyses in lambda grammars can be
preserved and/or corrected while the system keeps much of the flavor
of lambda grammars.
An alternative, especially for those convinced of the need to extend
lambda grammars to handle linguistic features and island constraints \cite{pp10acg,pompigne13}, is to move to
first-order linear logic, which also allows us to preserves the things
that work in lambda grammars but incorporate a simple treatment of
both features and island constraints without having to change the underlying logical theory. Those particularly attached to
dependent types can obtain them from MILL1 proofs by means of the
Curry-Howard isomorphism; first-order logic is a fairly weak fragment
of the lambda calculus with dependent types \cite{su06ch} so we need
to verify whether
it is expressive enough. Since smart parsing algorithms for lambda
grammars \cite{groote07parsing} already use first-order (linear) logic to
drive proof search, this
solution stays close to the computational core of lambda grammars: it
remedies the severe problems but also allows us to include treatments
for which much more complicated analyses have been proposed. However,
it is not clear in such a setup what the typed lambda terms actually
contribute and we would have a much simpler system if we simply
removed the typed lambda terms from the surface structure component
and handle all of the surface structure in first-order linear logic.
\section{Conclusions}
In this paper, we have shown that Hybrid Type-Logical Grammars
\cite{kl13coord} (and by extension lambda grammars/abstract categorial
grammars) can be embedded in first-order linear logic by means of a
simple translation, formula to formula and proof to proof. This
provides cleaner proof-theoretic foundations for Hybrid Type-Logical
Grammars but also suggests new ways of parsing these grammars. As an
immediate corollary, we have also shown that Hybrid Type-Logical
Grammars are NP-complete (like lambda grammars and the Lambek
calculus).
We have also seen how this translation provides a new perspective of
the known (but often ignored) problems of lambda grammars with
coordination and shown that the lack of left-right symmetry (or, at
the very
least, the absence of a way to emulate the Lambek calculus introduction
rules) results in overgeneration and descriptive inadequacy problems
for a much larger class of cases than previously assumed.
Combined with the results from \cite{mill1} and \cite{moot13lambek},
this means that the Lambek calculus, the Displacement calculus, lambda
grammars and Hybrid Type-Logical Grammars can all be translated into
first-order linear logic by means of simple translations and that,
moreover, many of the analyses of linguistic phenomena in these different
systems converge upon translation into first-order linear logic.
First-order linear logic can thus be seen as a way to decompose the
connectives of all these logics, separating the functor/argument
structure from the word order operations.
\section*{Acknowledgments}
This paper is deeply indebted to Yusuke Kubota and Robert Levine, whose ESSLLI 2013 course awoke my
curiosity both about the proof theoretic aspects of hybrid
type-logical grammar and about
the descriptive inadequacies of lambda grammars/abstract categorial
grammars --- the two principal themes of the current paper.
Early versions of these ideas were presented at the LIX Colloquium on
the Theory and Application of Formal Proofs (Palaiseau, November 2013), Computational
Linguistics in the Netherlands (Leiden, January 2014) and the Polymnie
workshop (Toulouse, March 2014). I would like all the people present
there for their questions and constructive comments, notably Crit Cremers, Philippe de Groote, Dominic Hughes
and
Dale Miller.
Last, but certainly not least, I would like to thank Michael Moortgat,
Carl Pollard and Christian Retor\'{e} for their discussion about the themes of this paper.
All remaining errors are of course my own.
This work has
benefitted from the generous
support of the French agency Agence Nationale de la Recherche as part of the project Polymnie
(ANR-12-CORD-0004).
\bibliographystyle{agsm}
|
\section{Introduction}
Two-dimensional electron gases (2DEGs) in transition-metal oxides (TMOs)
present remarkable phenomena that make them unique from a fundamental viewpoint
and promising for applications~\cite{Takagi2010,Mannhart2010}.
For instance, heterostructures grown on the $(001)$ surface of SrTiO$_3$,
a TMO insulator with a large band-gap of $\sim 3.5$~eV,
can develop 2DEGs showing metal-to-insulator transitions~\cite{Thiel2006},
superconductivity~\cite{Caviglia2008},
or magnetism~\cite{Brinkman2007,Salluzzo2013}.
Recently, 2DEGs at the $(111)$ and $(110)$ interfaces of LaAlO$_3$/SrTiO$_3$
were also reported~\cite{Herranz2012}.
The latter showed a highly anisotropic conductivity~\cite{Annadi2013}
and a superconducting state spatially more extended
than the one at the $(001)$ interface~\cite{Herranz2013}.
Interestingly, theoretical works have also predicted that exotic, possibly topological,
electronic states might occur at interfaces composed of $(111)$ bilayers
of cubic TMOs~\cite{Xiao2011,Yang2011,Ruegg2011,Doennig2013},
as two $(111)$ planes of transition-metal ions form a honeycomb lattice,
similar to the one found in graphene.
In this context, the discoveries that 2DEGs can also be created at the bare $(001)$ surfaces
of SrTiO$_3$~\cite{Santander-Syro2011,Meevasana2011,Plumb2013}
and KTaO$_3$~\cite{King2012,Santander-Syro2012},
and more recently at the $(111)$ surface of KTaO$_3$~\cite{Bareille2014},
opened new roads in the fabrication and study of different types of 2DEGs in TMOs
--in particular using surface-sensitive spectroscopic techniques,
which give direct information about the Fermi surface and subband structure
of the confined states.
The origin of the confinement is attributed to a local doping
of the surface region due to oxygen vacancies and/or lattice distortions.
Here we show that new types of 2DEGs can be directly tailored
at the bare $(110)$ and $(111)$ surfaces of SrTiO$_3$.
Imaging their electronic structure \emph{via} angle-resolved photoemission spectroscopy (ARPES),
we find that their Fermi surfaces, subband masses, and orbital ordering
are different from the ones of the 2DEG
at the SrTiO$_3$~$(001)$ surface~\cite{Santander-Syro2011,Meevasana2011}
and the ones predicted for the bulk,
being thus uniquely sensitive to the confining crystallographic direction.
This occurs because the crystallographic symmetries of the 2DEG plane,
and the electron effective masses along the confinement direction,
influence the symmetry of the electronic structure and the orbital ordering
of the $t_{2g}$ orbitals.
Furthermore, the observed carrier concentrations and 2DEG thicknesses
for different surfaces allow us to showcase the impact of oxygen vacancies
and of the polar discontinuity on distinctive features of the confined conducting sheet.
\begin{figure*
\begin{center}
\includegraphics[clip, width=14cm]{STO-111-VF-Fig1.eps}
\end{center}
\caption{\label{fig:fig1} \footnotesize
(a,~b) Unit cell of the cubic perovskite lattice of SrTiO$_3$.
The grey planes are the $(110)$ and $(111)$ planes, respectively.
The yellow dots represent the O$^{2-}$ anions, the black dot in the center
the Sr$^{2+}$ cation, and the red/green/blue dots the Ti$^{4+}$ cations
in different $(110)$ or $(111)$ planes.
Both orientations are highly polar,
as the crystal is built of alternating layers of (SrTiO)$^{4+}$ and
(O$_2)^{4-}$ or Ti$^{4+}$ and (SrO$_3)^{4-}$.
(c,~d) Ti$^{4+}$ cations of the crystal lattice at the $(110)$ and $(111)$ planes.
The black arrows indicate the lattice vectors of the Ti$^{4+}$ cations
in one $(110)$ or $(111)$ plane.
As indicated by the black lines in panel~(d),
a $(111)$-bilayer of Ti$^{4+}$ cations forms a honeycomb lattice.
(e) Bulk Fermi surface, calculated using a tight-binding model
with an unrealistically large value of $10^{21}$~cm$^{-3}$
for the bulk carrier density, intended to make the Fermi surface visible.
Such carrier density is at least \emph{three orders of magnitude} higher
than the bulk carrier density of the samples prepared for this study.
(f,~g) Cross section of the bulk Fermi surface in (e)
along the $(110)$ and $(111)$ planes, respectively.
The grey lines show the cross section of the bulk 3D Brillouin zone
through a $\Gamma$ point, while the black lines correspond
to the surface Brillouin zone.
}
}
\end{figure*}
The confined states were either created by fracturing the samples in vacuum
or by chemically and thermally preparing the surfaces \emph{in situ},
and studied through ARPES at the Synchrotron Radiation Center
(SRC, University of Wisconsin, Madison)
and the Synchrotron Soleil (France).
The sample preparation, similar to the one in references~\cite{Biswas2011,Chang2008},
is detailed in the Supplemental Material~\cite{Supplement}.
All through this paper, we describe the crystal structure in a cubic basis of unit-cell vectors,
and note as $[hkl]$ the crystallographic directions in real space,
$\langle hkl \rangle$ the corresponding directions in reciprocal space,
and as $(hkl)$ the planes orthogonal to those directions.
The major difference between the confined states
at various surface orientations of SrTiO$_3$
originates from the different symmetries of the corresponding crystal planes:
4-fold for the $(001)$ plane, 2-fold for the $(110)$ surface, and 6-fold for the $(111)$ surface.
Another difference is the polar character of the surface.
Thus, while the $(001)$ terminations, namely SrO or TiO$_2$, are nominally non-polar,
the $(110)$ terminations are alternatively $($SrTiO$)^{4+}$ and $($O$_2)^{4-}$,
and the $(111)$ terminations are either Ti$^{4+}$ or $($SrO$_3)^{4-}$.
These different surface symmetries and their polarity
are illustrated in figures~\ref{fig:fig1}(a-d).
Note in particular, from figure~\ref{fig:fig1}(d), that
a $(111)$-type bilayer of Ti$^{4+}$ cations forms a honeycomb lattice,
as noted in Ref.~\cite{Xiao2011}.
\begin{table}[b]
\caption{\label{Table:Masses} \footnotesize{
Effective light (L) and heavy (H) masses predicted by a TB model
in the bulk (first row) and experimental in-plane masses of the 2DEGs
at the $(001)$, $(110)$, and $(111)$ surfaces (other rows)
along the different high-symmetry directions of the crystal lattice (columns) of SrTiO$_{3}$.
In the bulk, all the effective masses along $\langle 111 \rangle$ are identical.
}
}
\begin{center}
\begin{tabular}{c | c c | c c | c c | c}
\hline \hline
& \multicolumn{2}{|c|}{$m_{100}/m_e$} & \multicolumn{2}{|c|}{$m_{110}/m_e$}
& \multicolumn{2}{|c|}{$m_{11\bar{2}}/m_e$} & $m_{111}/m_e$ \\
\hline
& $L$ & $H$ & $L$ & $H$ & $L$ & $H$ & \\
\hline
Theory bulk$^{a}$ & $1.06$ & $7.16$ & $1.06$ & $1.85$ & $1.24$ & $2.46$ & $1.48$ \\
\hline
SrTiO$_{3}$$(001)$ & $0.7^{b}$ & $10.0^{b}$ & $0.7^{c}$ & $1.3^{c}$
& $0.8^{c}$ & $1.8^{c}$ & $1.0^{c}$ \\
SrTiO$_{3}$$(110)$ & $1.0$ & $8.5$ & $1.6$ & $6.0$ & -- & -- & -- \\
SrTiO$_{3}$$(111)$ & -- & -- & $0.27$ & $1.08$ & $0.33$ & $8.67$ & -- \\
\hline \hline
\multicolumn{4}{l}{\footnotesize{$^{a}$ From Ref.~\cite{Khalsa2012}}} &
\multicolumn{4}{l}{\footnotesize{$^{b}$ From Ref.~\cite{Santander-Syro2011}}} \\
\multicolumn{8}{l}{\footnotesize{$^{c}$ From TB model using experimental masses
along $\langle 100 \rangle$}} \\
\end{tabular}
\end{center}
\end{table}
For our discussion later, it will be instructive to contrast the observations
at the $(110)$ and $(111)$ SrTiO$_3$ surfaces with both
the 2DEG at the $(001)$ surface and a model bulk electronic structure.
Figure~\ref{fig:fig1}(e) shows the bulk Fermi surface
from a simplified tight-binding (TB) model
where the electron hopping amplitudes between the three $t_{2g}$ orbitals
of neighboring Ti$^{4+}$ are $t_\pi=0.236$~eV and $t_{\delta'}=0.035$~eV~\cite{Khalsa2012},
and we neglect spin-orbit coupling and tetragonal distortions.
Near the $\Gamma$ point, this gives effectives masses listed in the first row
of table~\ref{Table:Masses} for various directions.
Figures~\ref{fig:fig1}(f,~g) show cross sections of the bulk Fermi surface
along the $(110)$ and $(111)$ planes through the $\Gamma$ point,
illustrating their respective 2-fold and 6-fold symmetries.
The \emph{experimental} spectra at the SrTiO$_3$ $(001)$ surface~\cite{Santander-Syro2011},
on the other hand,
fit well to a TB form where the hopping amplitudes are
$\bar{t}_\pi=0.36$~eV and $\bar{t}_{\delta'}=0.025$~eV,
leading to values of the effective masses near the $\Gamma$ point
shown in the second rows of table~\ref{Table:Masses}.
Note that all these masses differ by about 30\% from the bulk theoretical ones.
\begin{figure
\begin{center}
\includegraphics[clip, width=7cm]{STO-111-VF-Fig2.eps}
\end{center}
\caption{\label{fig:fig110} \footnotesize
(a)~ARPES Fermi surface map (second derivative) at $h\nu=91$~eV
in the $(110)$ plane of a fractured insulating SrTiO$_3$ sample.
The map is a superposition of intensities measured in the bulk $\Gamma_{130}$
and $\Gamma_{131}$ Brillouin zones~\cite{Supplement}.
The red lines indicate the edges of the unreconstructed $(110)$ Brillouin zones.
(b)~Energy-momentum intensity map at a $\Gamma$ point
along the $k_{\langle 001 \rangle}$ direction.
}
}
\end{figure}
We now present our experimental results.
Figure~\ref{fig:fig110}(a) shows the Fermi surface measured at the \emph{fractured} $(110)$ surface
of an undoped \emph{insulating} SrTiO$_3$ sample.
As we will see, our observations are similar
to another recent study of the 2DEG at the SrTiO$_3$(110) surface in a Nb-doped sample
prepared \textit{in situ} by Wang~\emph{et al.}~\cite{Wang2013}.
The metallic states we observe present the same 2-fold symmetry of the unreconstructed
$(110)$ surface Brillouin zone (BZ), represented by red rectangles.
This implies that \emph{(i)} the macroscopic properties of this 2DEG
should be highly anisotropic, echoing the observed anisotropic transport characteristics
reported in 2DEGs at $(110)$ LaAlO$_3$/SrTiO$_3$ interfaces~\cite{Annadi2013},
and \emph{(ii)} any surface roughness or reconstructions, expected in this highly polar surface,
do not affect the 2DEG, which must then reside in the sub-surface layers
--in agreement with our previous conclusions on fractured $(111)$ surfaces of KTaO$_3$~\cite{Bareille2014}.
Figure~\ref{fig:fig110}(b) shows the dispersion along the $k_{\langle 001 \rangle}$ direction,
giving rise to the longest of the two ellipsoidal Fermi surfaces in figure~\ref{fig:fig110}(a).
The band forming the shortest ellipsoid is eclipsed by photoemission selection rules
along this direction (see the Supplemental Material~\cite{Supplement}).
The band bottom and Fermi momenta are about $-40$~meV and $0.3$~\AA$^{-1}$, respectively.
From the data above, we model the Fermi surface of the 2DEG at the SrTiO$_3$~$(110)$ surface
as two orthogonal ellipses,
one along along $\langle 001 \rangle$ with semi-axes of 0.3~\AA$^{-1}$ and 0.1~\AA$^{-1}$,
the other along $\langle 1\bar{1}0 \rangle$ with semi-axes 0.25~\AA$^{-1}$ and 0.13~\AA$^{-1}$.
From the area $A_F$ enclosed by the Fermi surfaces, we obtain a carrier density
$n_{2D}^{(110)} = A_F/2\pi^2 \approx 1 \times 10^{14}$~cm$^{-2}$.
The electronic states associated to such a high charge carrier density
\emph{must be confined to the region near the surface}
--otherwise the bulk would be highly conductive,
in contradiction with the insulating nature of the samples studied.
Similarly, from the band bottom and Fermi momenta, using a parabolic approximation,
we obtain the effective band masses along $\langle 001 \rangle$
and $\langle \bar{1}\bar{1}0 \rangle$ (and equivalent directions),
listed in the third row of table~\ref{Table:Masses}.
These effective masses are similar to the ones determined in the aforementioned study~\cite{Wang2013}
of the 2DEG at the SrTiO$_3$(110) surface.
In our study, the band bottom of the heavy band, \emph{c.f.} figure \ref{fig:fig110}(b),
and the carrier density of the 2DEG are slightly lower,
probably due to the different surface preparation techniques.
Henceforth, we focus on new experimental results at the $(111)$ surface of SrTiO$_3$,
which as we will see presents the hexagonal symmetry of the unreconstructed surface,
and could thus be an interesting platform for the quest
of new electronic states and macroscopic properties at oxide surfaces.
\begin{figure
\begin{center}
\includegraphics[clip, width=7cm]{STO-111-VF-Fig3.eps}
\end{center}
\caption{\label{fig:fig111} \footnotesize
(a) Fermi surface map measured at $h\nu = 110$~eV
on a SrTiO$_3$ $(111)$ surface prepared \emph{in-situ}.
The black lines indicate the edges of the unreconstructed $(111)$
Brillouin zones around $\Gamma_{222}$.
(b) Fermi surface map (second derivative of ARPES intensity, negative values)
in the $k_{\langle 111 \rangle}$~--~$k_{\langle \bar{1}\bar{1}2 \rangle}$, or $(1\bar{1}0)$ plane,
acquired by measuring at normal emission while varying the photon energy in 1~eV steps
between $h\nu_1 = 67$~eV and $h\nu_2 = 120$~eV.
The experimental Fermi momenta, represented by the black and red circles,
were obtained by fitting the momentum distribution curves (MDCs)
integrated over $E_F \pm 5$~meV.
The red rectangle is the bulk Brillouin zone in the $(1\bar{1}0)$ plane.
(c) Energy-momentum map across the $\Gamma$ point along the $\langle \bar{1}\bar{1}2 \rangle$ direction.
The dispersions of a heavy band and light bands are visible.
(d)~Raw energy distribution curves of the dispersions shown in panel~(c).
In panels (a) and (c), the blue lines are simultaneous TB fits to the Fermi surface
and dispersions.
}
}
\end{figure}
Figure~\ref{fig:fig111}(a) shows the Fermi surface measured at the
SrTiO$_3$~$(111)$ surface prepared \emph{in-situ},
as described in the Supplemental Material~\cite{Supplement}.
It consists of three ellipses forming a six-pointed star,
thus strongly differing from the Fermi surface at the SrTiO$_3$~$(110)$ surface,
shown in figure~\ref{fig:fig110}(a), or the one at the SrTiO$_3$~$(001)$ surface,
discussed in previous works~\cite{Santander-Syro2011,Meevasana2011,Plumb2013}.
Additional experiments show that for surfaces prepared \emph{in-situ} with either
$(1\times 1)$ or $(3 \times 3)$ reconstructions,
the band structure and periodicity of the confined states are \emph{identical},
and correspond to the one expected from an \emph{unreconstructed surface}~\cite{Supplement}.
This indicates that the 2DEG at the SrTiO$_{3}$$(111)$ surface is also located in the sub-surface layers,
and is at best weakly affected by the surface reconstructions at the polar $(111)$ surface.
The 2D-like character of the electronic states is strictly demonstrated
from the Fermi surface map in the
$\langle 111 \rangle - \langle \bar{1}\bar{1}2 \rangle$ plane,
shown in figure~\ref{fig:fig111}(b).
Here, one sees that the bands do not disperse along $k_{\langle 111 \rangle}$
over more than half a bulk Brillouin zone,
thereby confirming the confined (\emph{i.e.}, localized) character of the electrons
along the $[111]$ direction in real space.
The modulation of the intensity in the Fermi surface map,
a typical feature of quantum well states~\cite{Mugarza2000,Hansen1999},
is discussed in the Supplemental Material~\cite{Supplement}.
Interestingly, note that the red rectangles
in figures~\ref{fig:fig110}(a) and~\ref{fig:fig111}(b)
represent the Brillouin zone in the $(110)$ (or equivalent) plane.
Yet, as seen from those figures, the shapes of the corresponding Fermi surfaces
are completely different. This directly shows the orientational tuning of the Fermi surface
due to different confinement directions.
Figure~\ref{fig:fig111}(c) shows the energy-momentum map at the $\Gamma$ point
along the $\langle \bar{1}\bar{1}2 \rangle$ direction,
corresponding to the major axis of the ellipsoids forming the 6-pointed-star Fermi surface.
The dispersions of one light band and one heavy band are clearly visible.
These constitute the ground state of the 2DEG.
Additional subbands are not observed, implying that the band bending at the surface
is too low to populate the upper quantum-well states.
Within our resolution, the heavy and light bands are degenerate at $\Gamma$,
with their band bottom located at about $-57$~meV.
We fit simultaneously these dispersions and the whole Fermi surface of figure~\ref{fig:fig111}(a)
using a simple tight-binding model~\cite{Supplement}.
The fit, shown by the continuous blue lines,
yields Fermi momenta of about $0.07$~\AA$^{-1}$ and $0.36$~\AA$^{-1}$
for, respectively, the light and heavy bands along $\langle \bar{1}\bar{1}2 \rangle$.
This gives an electron concentration
$n_{2D}^{(111)} \approx 1.0 \times 10^{14}$~cm$^{-2}$,
and effective masses listed in the third row of table~\ref{Table:Masses}.
We now draw some comparisons between the effective masses
and thicknesses of the 2DEGs at the SrTiO$_3$~$(001)$, $(110)$ and $(111)$ surfaces.
Table~\ref{Table:Masses} shows that,
while the masses along the ``natural" electron-hopping directions in the bulk
($[001]$ and equivalent) are comparable between the 2DEGs
at the SrTiO$_3$~$(001)$ and $(110)$ surfaces, the masses along $[110]$ at the $(110)$ surface,
and all the masses of the 2DEG at the $(111)$ surface,
are very different from the ones expected from the tight-binding parameters
describing the bulk or the 2DEG at the $(001)$ surface.
In this respect,
note that if the confinement direction is $[110]$ or $[111]$,
then the electrons moving in the 2DEG plane along a direction \emph{other than $[001]$}
will experience the confining potential gradient and the modified crystal field outside the surface,
as they will hop in staircase patterns between first neighbors along $[001]$
(or equivalent) directions --see figures~\ref{fig:fig1}(a-d) and Ref.~\cite{Annadi2013,Bareille2014}.
The understanding of these mass differences,
also reported in quantum well states at thin films of simple-metals~\cite{Wu2002}
or strongly-correlated oxides~\cite{Yoshimatsu2011},
should be the subject of further theoretical works.
The maximal spatial extension $d_{max}$ of the 2DEGs at the SrTiO$_3$ $(110)$ and $(111)$ surfaces
can be estimated using a triangular potential well model~\cite{Supplement}.
We obtain $d_{max}^{110} \approx 1.7$~nm, which amounts to 6 2D-layers
or 3 bulk unit cells along $[110]$,
and $d_{max}^{111} \approx 1.9$~nm, corresponding to $\sim 9$ layers of Ti~$(111)$,
or again about 3 bulk unit cells along $[111]$.
Finally, we note that the orbital ordering of the electronic states
at the $(110)$ and $(111)$ surfaces of SrTiO$_{3}$ is different from the one at the $(100)$ surface.
In the first two cases, the bands are degenerate within our experimental resolution, whereas
at the $(001)$ surface the smallest observed splitting between bands
of different orbital character is of 50~meV~\cite{Santander-Syro2011}.
As the confinement energy of each band is inversely proportional
to its effective mass along the confinement direction~\cite{Santander-Syro2011},
different surface orientations result in different orbital ordering.
But along the $[111]$ direction the effective masses of the three $t_{2g}$ bands
are identical, and so their degeneracy at the $\Gamma$ point is not lifted by the confinement.
Similarly, the effective masses of bands of different orbital character
along $[110]$ are quite similar (see table~\ref{Table:Masses}).
Hence, the degeneracy lift is rather small, and cannot be observed in our data.
This demonstrates the influence of the confinement direction on the orbital ordering.
Several scenarios have been proposed to explain the origin of the 2DEG
at the LaAlO$_3$/SrTiO$_3$ $(001)$ interface.
According to one of these, the formation of a conducting sheet
prevents the occurrence of a polar catastrophe in the material.
Yet, the discovery of a confined 2DEG at the $(001)$ surface of SrTiO$_3$,
with characteristics similar to those of the above heterostructure,
suggests that the driving mechanism may not be unique,
as in the bare SrTiO$_3$ all the layers are electrically neutral.
Instead, in the latter case, surface oxygen vacancies are believed to cause
and to confine the gas~\cite{Santander-Syro2011,Meevasana2011,Wang2013}.
Additionally, for the $(110)$ and $(111)$ SrTiO$_3$ surfaces,
of nominal polar charge $4e$,
one would expect a much larger carrier concentration in the 2DEG,
and a very strong electric field confining the electrons in a narrow sheet at the surface.
However, we observe that the carrier concentrations and thicknesses of the 2DEGs
are quite comparable for all three orientations (this work and Ref.~\cite{Santander-Syro2011}):
$n_{2D} \sim 10^{14}$~cm$^{-2}$, $d_{max} \sim 2$~nm.
In fact, in the polar SrTiO$_3$ surfaces studied here,
the polar catastrophe does not seem to be compensated
by the electrons of the 2DEG but by surface reconstructions or relaxations,
while the 2DEG lies in the subsurface layers.
Thus, although the 2D electronic structure
(effective masses, orbital ordering)
depends on the surface orientation,
the thickness and carrier concentration of the 2DEG might be controlled by another factor,
probably oxygen vacancies
and/or lattice distortions induced
by the synchrotron light irradiation, as discussed in the Supplemental Material~\cite{Supplement}.
In conclusion, our results show that the symmetries, electronic structure,
and orbital ordering of the confined states at the surface of TMOs
can be tailored by confining the electrons along different directions
in the \emph{same} material.
Such orientational tuning echoes the differences of transport properties
reported recently in LaAlO$_3$/SrTiO$_3$ $(110)$ and $(111)$
interfaces~\cite{Herranz2012,Annadi2013,Herranz2013}.
In particular, from our data, the highly anisotropic transport behavior
observed in the $(110)$ interfaces~\cite{Annadi2013}
can be directly related to the 2-fold symmetry of the Fermi surface measured by ARPES.
More generally, our results provide an exciting route for obtaining
new types of 2D electronic states in correlated-electron oxides.
We thank V. Pillard for her contribution to the sample preparation.
T.C.R. acknowledges funding from the RTRA Triangle de la Physique (project PEGASOS).
A.F.S.-S. and M.G. acknowledge support from the Institut Universitaire de France.
This work is supported by public grants from the French National Research Agency (ANR)
(project LACUNES No ANR-13-BS04-0006-01)
and the ``Laboratoire d'Excellence Physique Atomes Lumi\`ere Mati\`ere'' (LabEx PALM project ELECTROX)
overseen by the ANR as part of the ``Investissements d'Avenir'' program (reference: ANR-10-LABX-0039).
\section{Supplemental Material}
\subsection*{ARPES Experiments}
The ARPES measurements were conducted at the Synchrotron Radiation Center
(SRC, University of Wisconsin, Madison)
and the Synchrotron Soleil (France).
We used linearly polarized photons in the energy range $20-120$~eV,
and Scienta R4000 electron detectors with vertical slits.
The angle and energy resolutions were $0.25^{\circ}$ and 25~meV at SRC,
and $0.25^{\circ}$ and 15~meV at Soleil.
The mean diameter of the incident photon beam was smaller than 100~$\mu$m.
The samples were cooled down to 10-30~K before fracturing or measuring,
in pressure lower than $6\times10^{-11}$~Torr.
The confined states were either created by fracturing the samples in vacuum
or by chemically and thermally preparing the surfaces \emph{in situ},
as detailed in the next section.
The results were reproduced for at least five different samples for each surface orientation.
\subsection*{Surface preparation}
\begin{figure*}
\begin{center}
\includegraphics[clip, width=14cm]{STO-111-VF-FigS1.eps}
\end{center}
\caption{\label{fig:sto111_afm} \footnotesize{(Color online)
(a) Atomic force microscope (AFM) image of a chemically
and thermally prepared SrTiO$_3$~$(111)$ surface.
The surface is single terminated and unreconstructed,
as shown in the RHEED image in (b).
Longer annealing times result in a mixed terminated surface,
as demonstrated in the AFM friction image (c) measured in contact mode.
A $3 \times 3$ reconstruction of the surface can be deduced
from the corresponding RHEED image in (d).
}
}
\end{figure*}
The non-doped, polished crystals of SrTiO$_3$ were supplied by CrysTec GmbH and Aldrich.
To prepare the surface, the samples were ultrasonically agitated in deionized water,
subsequently etched in buffered HF and annealed at $950^{\circ}$C
for several hours in oxygen flow.
Depending on the annealing time, this treatment yields a Ti-rich,
single-terminated or mixed-terminated step-and-terrace structured
surface of SrTiO$_3$~$(111)$~\cite{Biswas2011}.
Figure~\ref{fig:sto111_afm}(a) shows the atomic-force microscopy (AFM) image
of the single-terminated $(111)$ surface of a sample annealed for 3h.
This treatment produces a $(1 \times 1)$ unreconstructed surface,
shown by RHEED image in figure~\ref{fig:sto111_afm}(b).
Longer annealing (10h) results in a mixed-terminated surface~\cite{Chang2008},
as shown in the AFM friction image in figure~\ref{fig:sto111_afm}(c),
measured in contact mode.
The surface prepared in such a way is $(3 \times 3)$ reconstructed,
as displayed in the RHEED image in figure~\ref{fig:sto111_afm}(d).
The surface state of the cleaved samples
was not determined by imaging or diffraction techniques.
To perform the surface-sensitive ARPES measurements,
one needs pristine and crystalline surfaces.
To clean the surface of contaminations, the samples prepared as described above
were further annealed \emph{in-situ} in vacuum at a pressure of
approximately $p=3\times10^{-9}$~mbar at a temperature of $T = 550^\circ$C for about 2 hours.
This annealing step cleans the surface, does not change the surface reconstruction,
and also introduces oxygen vacancies in the bulk of the SrTiO$_3$ samples.
Note that the introduced bulk charge carrier density is at least
three orders of magnitude lower than the one observed for the confined states
in the ARPES measurements, as detailed in the main text.
Moreover, Plumb~\emph{et al.} demonstrated that various
\emph{in-situ} sample preparations, including annealing in an O$_2$-rich atmosphere
which results in a non-doped bulk, create identical confined states
at the $(001)$ surface of TiO$_2$-terminated SrTiO$_3$ ~\cite{Plumb2013}.
Recall also, from figure~\ref{fig:fig111}(b), that the states observed in our experiments
do not disperse along the confinement direction, which demonstrates
their quasi-2D character.
For the confined states at the $(111)$ surface, the quality of the obtained ARPES data
is better for the surface prepared \emph{in-situ}.
This might be due to the strong polar nature of the $(111)$ surface of SrTiO$_3$.
Hence, fracturing a sample along a $(111)$ plane might yield a partly disordered surface.
The electronic structure of the 2DEG at the SrTiO$_3$~$(111)$ surface
is similar for the cleaved and the two differently prepared surfaces
(unreconstructed and $(3 \times 3)$ reconstructed).
In fact, for all three types of surfaces the periodicity of the electronic structure
in reciprocal space, shown in figure~\ref{fig:fig_sup}(a) for the prepared, $(3 \times 3)$
reconstructed surface, corresponds to the one expected of an unreconstructed surface.
By Bloch theorem, the very existence of dispersive bands and well-defined Fermi surfaces
implies the existence of a periodic in-plane potential acting on the confined electrons,
hence of crystalline order at the layer(s) where the 2DEG is located.
As the electronic structure has the periodicity of the \emph{unreconstructed} surface,
the 2DEG seems to stabilize in a sub-surface region, where it is not affected by
any surface reconstructions or superstructures related to vicinal surfaces or terraces.
A possible explanation for this observation would be that the electrons of the Ti cations
in the topmost layer are localized, while the itinerant electrons exist in the subsurface layers.
For the $(110)$ surface, a surface preparation similar to the one described above
for the $(111)$ surface was conducted.
The data quality of fractured and prepared samples are quite similar
as the chemical etching step is not perfectly adapted to the $(110)$ surface.
Sr and Ti are both situated in one of the alternating $(110)$ layers
of $($SrTiO$)^{4+}$ and O$_2^{4-}$ building up the crystal lattice.
Thus, the selective etching of Sr-related species might result in a rather rough surface.
\subsection*{Photon energy dependence}
\begin{figure}
\begin{center}
\includegraphics[clip, width=7cm]{STO-111-VF-FigS2.eps}
\end{center}
\caption{\label{fig:fig_sup} \footnotesize{(Color online)
(a) Superposition of Fermi surface maps measured
for the chemically and thermally prepared SrTiO$_3$~$(111)$ sample
($(3\times 3)$ reconstructed surface)
at photon energies of $h\nu = 47$~eV and $h\nu=96$~eV.
(b) Reciprocal 2D space in the $(111)$ plane.
Inside each Brillouin zone the projections of the different bulk $\Gamma$ points
corresponding to available final states during the photoemission
process at the specified photon energy are indicated.
This diagram helps understanding the Fermi-surface intensities shown in panel (a).
The color (red, blue, green) of the hexagons
indicates which $\Gamma$ points are located in the same (111) plane in reciprocal space.
}
}
\end{figure}
The photon energy dependence of the electronic states at the SrTiO$_3$~$(111)$ surface
is displayed in the main text in figure~\ref{fig:fig111}.
Although the states do not disperse, confirming their confined nature,
the intensity of the states drops rather quickly moving away from $\Gamma_{222}$.
This observation is similar to the intensity modulation as a function of the photon energy
reported previously at the $(001)$ surface of SrTiO$_3$~\cite{Santander-Syro2011}
and KTaO$_3$~\cite{Santander-Syro2012}, as well as in quantum well states
of metals~\cite{Mugarza2000,Hansen1999}.
This modulation is due to photoemission dipole selection rules:
the optical excitation of the electrons occurs from initial states
in the near surface region that do not disperse along the confinement direction
(the confined electrons) to dispersing bulk final states.
Moreover, if the wave function of the confined states is not exactly
localized in a 2D layer, but exists over several unit cells,
the dispersion along the confinement direction will be affected.
This can be intuitively understood from Heisenberg uncertainty principle:
only a strict 2D confinement in real space yields a complete indetermination
of the electron momentum along the confinement direction,
hence an exactly cylindrical Fermi surface.
Some delocalization along the confinement direction, as in quantum-well states,
implies a small dispersion
of the Fermi surface along that direction.
Bearing these effects (selection rules in quantum wells, finite delocalization) in mind,
one can comprehend the data in figure~\ref{fig:fig_sup}(a),
which shows a superposition of Fermi surface maps measured at different photon energies,
for a $(111)$ surface prepared \emph{in-situ}.
The black hexagons are the Brillouin zones
assuming an unreconstructed surface.
Thus, due to selection rules, the intensity of the photoemission peak
from the confined states is highest close to positions
corresponding to $\Gamma$ points of the bulk,
where final states at the same $k_{\langle 111 \rangle}$ momentum are available for the optical transition.
But this intensity will decrease rapidly by moving along $k_{\langle 111 \rangle}$,
away from the bulk $\Gamma$ points~\cite{Mugarza2000}.
Experimentally, this is done by changing the photon energy.
This results in the necessity to measure in-plane Fermi surface maps
at different photon energies, and then superpose them
to retrieve the complete periodicity of the electronic states,
as illustrated in figure~\ref{fig:fig_sup}(b). This figure shows the positions
of the experimentally observed $\Gamma$ points projected in the $(111)$ plane.
The photon energy inside each Brillouin zone
corresponds to the $k_{\bot}$ value of the $\Gamma$ points
assuming a work function of $W = 4.25$~eV and an inner potential of $V_{0} = 12$~eV.
\subsection*{Fermi surface of S\lowercase{r}T\lowercase{i}O$_3$(110)}
\begin{figure}
\begin{center}
\includegraphics[clip, width=5cm]{STO-111-VF-FigS3.eps}
\end{center}
\caption{\label{fig:sto110_unfoldedFS} \footnotesize{(Color online)
Second derivative of ARPES Fermi surface map at $h\nu=91$~eV
in the $(110)$ plane of a cleaved insulating SrTiO$_3$ sample.
The map spans the $\Gamma_{130}$ (bottom)
and $\Gamma_{131}$ (top) Brillouin zones.
The red lines indicate the edges of the unreconstructed $(110)$ Brillouin zones.
}
}
\end{figure}
As stated in the main text,
the Fermi surface map shown in figure~\ref{fig:fig110}
is a superposition of intensities measured in the bulk $\Gamma_{130}$ and $\Gamma_{131}$ Brillouin zones.
Figure~\ref{fig:sto110_unfoldedFS} shows the intensities measured
in those Brillouin zones.
Due to photoemission matrix elements, only the vertical ellipsoidal Fermi surface
is observed around the $\Gamma_{130}$ point,
while both the vertical and the smaller horizontal ellipsoidal Fermi surfaces
are observed around the $\Gamma_{131}$ point.
\subsection*{Estimate of the spatial extensions of the 2DEGs at the S\lowercase{r}T\lowercase{i}O$_3$~$(110)$ and $(111)$ surfaces.}
In our data, figures~2 and~3 of the main text, only the lowest-energy subbands are observed.
To estimate the maximal extension $d_{max}$ of the corresponding confined states,
we follow the same strategy of Ref.~\cite{Bareille2014}.
We assume that the second subbands are slightly above the Fermi level,
hence unoccupied and not detectable by ARPES.
We then use a triangular potential well model, and take as effective masses
along the $[110]$ and $[111]$ confinement directions, respectively,
$m_{110} \approx 1.6 m_e$
(the lightest of the masses gives the largest 2DEG thickness)
and $m_{111} = 1.0 m_e$ (given by extrapolating the experimental masses at the $(001)$ surface
to the bulk $[111]$ direction) --see table~I of the main text.
This gives $d_{max}^{110} \approx 1.7$~nm, amounting to 6 2D-layers
or 3 bulk unit cells along $[110]$,
and $d_{max}^{111} \approx 1.9$~nm, corresponding to $\sim 9$ layers of Ti~$(111)$,
or again about 3 bulk unit cells along $[111]$.
\subsection*{UV dose dependence: enhancement of T\lowercase{i}$^{3+}$ signal}
\begin{figure}
\begin{center}
\includegraphics[clip, width=8cm]{STO-111-VF-FigS4.eps}
\end{center}
\caption{\label{fig:fig_UVdose} \footnotesize{(Color online)
(a) Angle-integrated spectra of an SrTiO$_3$ sample prepared \emph{in-situ},
measured at a photon energy of $hv=110$~eV, with a step size of 50~meV,
showing the density of states for binding energies between $-45$~eV and $2$~eV.
The black curve was measured shortly after the first exposure of the sample
to the UV light, and the red curve at the end of the measurements (about 36 hours later).
(b) Zoom over the valence band region.
(c) Angle-integrated spectra showing the in-gap states
and the confined states at the Fermi level,
measured at $hv=110$~eV with a step size of 5meV.
(d) Zoom over the confined states at the Fermi level.
}
}
\end{figure}
Understanding the influence of the UV synchrotron illumination
on the observed confined states is important to determine the origin of such states.
Recent photoemission studies on the 2DEGs
at the $(001)$ or $(110)$ surface of SrTiO$_3$
proposed that the UV light creates oxygen vacancies~\cite{Meevasana2011,Wang2013} or, respectively,
ferroelectric lattice distortions~\cite{Plumb2013} in the surface region.
The two effects are difficult to disentangle using photoemission,
as in both cases charge is transferred from O to Ti.
Figure\ref{fig:fig_UVdose}(a) shows the angle-integrated spectra,
measured at $hv=110$~eV, of a SrTiO$_3$ sample prepared \emph{in-situ}
for binding energies between $-45$~eV and $2$~eV.
The black curve was measured
shortly after the first exposure of the sample to the UV light,
while the red curve was recorded at the end of the measurements (36 hours later).
The spectra are normalized to the intensity of the Sr $4p$-peak,
which should be rather independent of the concentration of oxygen vacancies and/or
ferroelectric lattice distortions.
Figure~\ref{fig:fig_UVdose}(b) is a zoom over the valence band region,
while figures~\ref{fig:fig_UVdose}(c,~d) show the in-gap states and the
confined states at the Fermi level.
The change of various features under UV irradiation is obvious:
first, the formation of a shoulder in the Ti-$3p$ peak
at lower binding energies, indicating electron transfer from Ti$^{4+}$ to a lower valency state.
Second, the decrease in intensity of the valence band in its low binding energy region.
Third, the increase in intensity of the in-gap states
and of the peak corresponding to the confined states.
All these observations could be explained by both scenarios:
the creation of oxygen vacancies and the ferroelectric lattice distortions.
In contrast to samples prepared \emph{in-situ},
cleaved samples show a different behavior regarding the UV light exposure.
The subbands of the 2DEG in all the \emph{cleaved} SrTiO$_3$ surfaces
we have studied so far, \emph{i.e.} $(001)$, $(110)$ and $(111)$,
are all observed essentially \emph{immediately after cleaving},
with no or little time delay after the first exposure to UV light.
A more detailed study on the UV induced effects is beyond the scope of this paper.
\subsection*{Tight-binding calculations of the 2DEG at the S\lowercase{r}T\lowercase{i}O$_3$$(111)$ surface}
The band dispersions shown in the main text correspond
to the bottom of the conduction band of SrTiO$_3$,
which is formed by Ti-$3d$ orbitals hybridized with O-$2p$ orbitals.
The interaction between the oxygen anions forming an octahedron
and the Sr cation generates a large crystal field which splits the $d$ states
in a lower $t_{2g}$ triplet and an higher $e_{g}$ doublet.
Hence, only the $t_{2g}$ orbitals are considered in our tight-binding model,
which is based on the calculations of reference~\cite{Xiao2011}.
Our model for the SrTiO$_3$$(111)$ surface is limited to a bilayer of Ti atoms.
This approach is sufficient to fit the experimental data as shown in the main text,
but does not necessarily imply the confinement of the electrons to a bilayer.
The Hamiltonian $H$ of the system in the basis $\{d_{I,n}\}$,
where $I=(X,Y,Z)$ correspond to the orbital character $(yz,zx,xy)$ of the $t_{2g}$ orbitals
and $n=1,2$ indicates the number of the layer of Ti cations, is given by:
\begin{equation*}
H=
\begin{pmatrix}
d^\dagger_{X,1} \\
d^\dagger_{Y,1} \\
d^\dagger_{Z,1} \\
d^\dagger_{X,2} \\
d^\dagger_{Y,2} \\
d^\dagger_{Z,2}
\end{pmatrix}
^T
\begin{pmatrix}
\tilde{\epsilon}_X & & & \epsilon_{X} & & \\
& \tilde{\epsilon}_Y & & & \epsilon_{Y} & \\
& & \tilde{\epsilon}_Z & & & \epsilon_{Z} \\
\epsilon_X^* & & & \tilde{\epsilon}_{X} & & \\
& \epsilon_Y^* & & & \tilde{\epsilon}_{Y} & \\
& & \epsilon_Z^* & & & \tilde{\epsilon}_{Z} \\
\end{pmatrix}
\begin{pmatrix}
d_{X,1} \\
d_{Y,1} \\
d_{Z,1} \\
d_{X,2} \\
d_{Y,2} \\
d_{Z,2}.
\end{pmatrix}
\end{equation*}
Here, $\tilde{\epsilon}_I$ describes the hopping of electrons
between next nearest neighbors of Ti cations (intra-layer hopping),
characterized by the hopping amplitude $t_{\sigma''}$,
whereas $\epsilon_I$ describes the hopping between nearest neighbors
(inter-layer hopping) with hopping amplitudes $t_\pi$ and $t{_\delta'}$:
\begin{align*}
\tilde{\epsilon}_X&=-2t_{\sigma''}\cos(-\frac{\sqrt{3}}{2}\tilde{a}~k_x + \frac{3}{2}b~k_y)\\
\tilde{\epsilon}_Y&=-2t_{\sigma''}\cos(\frac{\sqrt{3}}{2}\tilde{a}~k_x + \frac{3}{2}b~k_y)\\
\tilde{\epsilon}_Z&=-2t_{\sigma''}\cos(\sqrt{3}\tilde{a}~k_x)\\
\epsilon_X&=-t_\pi e^{-i\tilde{a}~k_y}
\left[1+e^{i\frac{\tilde{a}}{2}(-\sqrt{3}k_x+3k_y)}\right]
-t_{\delta'}e^{i\frac{\tilde{a}}{2}(\sqrt{3}k_x+k_y)}\\
\epsilon_Y&=-t_\pi e^{-i\tilde{a}~k_y}
\left[1+e^{i\frac{\tilde{a}}{2}(\sqrt{3}k_x+3k_y)}\right]
-t_{\delta'}e^{i\frac{\tilde{a}}{2}(-\sqrt{3}k_x+k_y)}\\
\epsilon_Z&=-2t_\pi e^{\frac{i}{2}\tilde{a}~k_y}
\cos(\frac{\sqrt{3}}{2}\tilde{a}~k_x)-t_{\delta'} e^{-i\tilde{a}~k_y}.
\end{align*}
In the above expressions, $k_x$ corresponds to $k_{\langle 1\bar{1}0 \rangle}$,
$k_y$ to $k_{\bar{1}\bar{1}2}$, and $\tilde{a}$ to the cubic lattice constant $a$
projected in the (111) plane $\tilde{a}=\sqrt{2/3}a$.
Compared to the calculations of reference~\cite{Xiao2011}, our data can be fitted rather well
using a simplified model. We neglect in our model the spin-orbit coupling,
the trigonal crystal field, the layer potential difference,
crystal distortions at low temperature, and the hopping ($t_{\pi'}$)
between next nearest neighbors of different orbital symmetry.
The fits shown in figures~\ref{fig:fig111}(a) and \ref{fig:fig111}(c) of the main text
are based on such a simplified model using fitting parameters of
$t_\pi=1.6$~eV, $t_{\delta'}=0.07$~eV and $t_{\sigma''}=0.05$~eV.
Note that such value of $t_\pi$, which quantifies the hopping energy between nearest neighbors
along the $[100]$ (and equivalent) directions, is here over 4 times larger than
the same parameter inferred from the 2DEG at the SrTiO$_3$~$(001)$ surface
(namely, $t=0.36$~eV, see the main text). This shows again that the effective masses
of the 2DEG at the SrTiO$_3$~$(111)$ surface strongly differ from what would be
expected from a model based on the 2DEG at the $(001)$ surface.
As discussed in the main text, the electrons moving along any direction in the $(111)$ plane
will actually hop in zig-zag patterns between first neighbors along $[001]$
(or equivalent) directions, and thus will experience the confining potential gradient
and the modified crystal field outside the surface. These effects are not accounted by
our minimalist TB model.
Additionally, our TB model only considers one bilayer of Ti atoms.
However, it is known that in quantum well states the effective masses of the confined electrons
depend on the width of the quantum well or, equivalently,
the number of layers~\cite{Wu2002,Yoshimatsu2011}.
All these effects should be taken into account in future theoretical works addressing the 2DEGs at
the different surfaces of SrTiO$_3$.
On the other hand, while distortions of the crystal lattice,
and thereby of the overlap between the different $t_{2g}$ orbitals,
might exist at the surface and be slightly different depending on the surface orientations,
they should bear a negligible effect on the 2DEGs reported here, as we have seen that
their electronic structure is essentially insensitive to surface polarity or reconstructions.
|
\section{Contraction properties and eigenfunctions for Products of Positive Matrices}
\section{Introduction}
Let $d \ge 1$, $\abs{\cdot}$ be any norm on $\R^d$ and $\norm{\cdot}$ be the corresponding operator norm. Let $(\mA_n)_{n \in \N}$ be a sequence of independent identically distributed $d \times d$-matrices such that $\E \log^+ \norm{\mA_1} < \infty$. The Furstenberg-Kesten theorem {\color{black} \cite[Theorem 2]{Furstenberg1960}} provides us with a strong law of large numbers for the norm of the products $\mPi_n:= \mA_n \cdots \mA_1$, namely
$$ \lim_{n \to \infty}\frac{1}{n} \log \norm{ \mPi_n } = \gamma \qquad \Pfs,$$
with $\gamma = \inf_{m \in \N} m^{-1} \E\, {\color{black} \log} \norm{\mPi_m}$ being called the (top) Lyapunov exponent of $(\mA_n)_{n \in \N}$.
Under different sets of additional assumptions (to be detailed below) on the law $\mu$ of $A_1$, the convergence result has been strengthened towards a SLLN for the norm of a vector under the action of the random matrices: For example, following \cite{Cohn1993,Hennion1997,Kesten1973,Kingman1973}, assume that the support of $\mu$ consists of nonnegative matrices and contains a matrix with all entries positive. Then it holds for all nonnegative vectors $x$ that
\begin{equation}\label{furstenbergkesten} \lim_{n \to \infty}\frac{1}{n} S_n^x := \lim_{n \to \infty}\frac{1}{n} \log \abs{\mPi_n x} = \gamma \qquad \Pfs \end{equation}
Under a second moment assumption, Hennion \cite{Hennion1997} proved a CLT, namely that
$$ \frac{1}{\sqrt{n}} \left(S_n^x - n \gamma\right)$$ converges to a normal law. For related limit theorems for invertible matrices, see \cite{Bougerol1985,LePage1982}.
Observe that in both cases, the SLLN and the CLT, the limit does not depend on the starting vector $x$. In contrast therewith is the result of Kesten \cite{Kesten1973} about the behavior of the maximum of $S_n^x$: Assuming in essence that the action of $\mA_1$ is both expanding and contracting with positive probability, that is $\gamma < 0$ but $\P{S_1^x > 0}>0$, Kesten showed that there is $\alpha > 0$ and a continuous function $\es[]$ on the unit sphere $\Sd$, which is strictly positive on nonnegative vectors, such that
\begin{equation}\label{eq:Kesten max} \lim_{t \to \infty} e^{\alpha t}\, \P{\max_{n} S_n^x > t } = \es[](x).
\end{equation}
Here the behavior in the limit depends on the initial value.
We are going to provide a third-order Edgeworth expansion, which gives a rate of convergence for the CLT. We also provide a formula for the asymptotic variance $\sigma^2$ and show that it is positive under a natural nonlattice assumption. The Edgeworth expansion will as well be the main tool in describing the convergence in the law of large numbers, i.e. the Furstenberg-Kesten theorem, in more details. In particular, we will discover how fluctuations depend on the starting vector $x$ as well as on the action of $(\mA_n)_{n \in \N}$ on the unit sphere, which is given by the Markov chain
$$ X_n^x := \frac{\mPi_n x}{\abs{\mPi_n x}}.$$
What we will prove is a large deviation result similar to the Bahadur-Rao theorem, i.e. for (suitable) $q > \gamma$, there is an explicitly given sequence $J_n(q)$ tending to infinity at an exponential rate, such that
\begin{equation} \label{eq:adhocBahadurRao} \lim_{n \to \infty} J_n(q) \Erw{r_q(X_n^x)\1[{\{{S_n^x} \ge nq\}}]} = r_q(x) \end{equation}
for a positive continuous function $r_q$ which depends on $q$, and generalizes the function $\es[](x)$ of Kesten's result.
This result is in the scope of large deviation principles for Markov additive processes, see \cite{Iltis2000,Kontoyiannis2003,Ney1987} for related results, where stronger conditions on $X_n^x$ have to be imposed than those who are satisfied for the chain generated by matrices. The very recent paper of Guivarc'h \cite{Guivarch2014} provides a local limit theorem, which is proved along similar lines as our Edgeworth expansion.
As an application of our result, we will shed new light on the classical result of Kesten about random difference equations: Let {\color{black}$\mM$} be a random $d \times d$-matrix and $B$ a random vector in $\R^d$. Under weak assumptions on {\color{black}$(\mM,B)$}, there is a unique solution (in law) to the equation
\begin{equation}\label{eq:rde1} R \eqdist {\color{black}\mM} R + B,\end{equation}
where $\eqdist$ means same law.
In the case of nonnegative ${\color{black} \mM}, B, R$, Kesten \cite{Kesten1973} proved, assuming that ${\color{black} \mM}$ is both contracting and expanding with positive probability, that for the same $\alpha > 0$ and $\es[]$ as in \eqref{eq:Kesten max},
\begin{equation}\label{eq:tails} \lim_{t \to \infty} \, t^\alpha \, \P{\skalar{R,x} > t} = K \es[](x), \end{equation}
for some $K >0$. This result has been extended to the case of invertible matrices in \cite{AM2010,BDGHU2009, Guivarch2012,Klueppelberg2004,LePage1983}, where it has always been an involved question to prove that $K$ is actually positive.
In both cases (nonnegative resp. invertible matrices), our result will be applied to give an rather elementary proof of the fact that $K > 0$. {\color{black} Here, the law of the matrix $\mA := \mM^\top$ will be relevant.}
This approach can also be extended to the study of branching equations, i.e.
\begin{equation}\label{eq:branching} R \eqdist \sum_{i=1}^N {\color{black} \mM_i} R_i + B, \end{equation}
where now $N \ge 2$ is a fixed integer, {\color{black} $(\mM_1, \dots, \mM_N)$} are random matrices and $B$ a random vector, independent of $R_i$, which are i.i.d. copies of $R$. For random variables $R$ satisfying such an equation, the heavy tail property \eqref{eq:tails} has been shown to hold in \cite{BDGM2014,BDMM2013,Mirek2013}, but the positivity of $K$ remained a partially open question in the latter two articles. Due to the branching structure of Eq. \eqref{eq:branching}, the combinatorial part of the approach becomes more involved (it has been worked out in the one-dimensional case in \cite{BDZ2014+}), this is why we decided to postpone it to the separate work \cite{BM2015} and focus on the application of the large deviations result here, which can be seen more directly in the case of Eq. \eqref{eq:rde1}.
Having thus described the scope of the paper, we are now going to introduce some notations and concepts in order to state the main results in full detail. Since we want to solve questions concerned with nonnegative matrices as well as with invertible matrices, we are led to introduce several sets of assumptions (namely those of Kesten \cite{Kesten1973}, Guivarc'h and Le Page \cite{Guivarch2012} and Alsmeyer and Mentemeier \cite{AM2010}) on the law $\mu$ of the random matrix $\mA$, with all of them being sufficient for the announced results to hold. The main focus will be on nonnegative matrices, for which we will provide details of proofs, while for invertible matrices, we will mainly highlight the differences and refer to the works cited above.
\section{Notations and Preliminaries}
We start by introducing three sets of assumptions for random matrices. Let $d \ge 1$. Given a probability law $\mu$ on the set of $d \times d$-matrices $M(d \times d, \R)$, let $(\mA_n)_{n \in \N}$ be a sequence of i.i.d.~random matrices with law $\mu$ and write $\mPi_n:= \mA_n \cdots \mA_1$ for the $n$-fold product. Equip $\Rd$ with any norm $\abs{\cdot}$, write $\norm{\ma}:=\sup_{x \in \Sd} \abs{\ma x}$ for the operator norm of a matrix $\ma$ and denote the unit sphere in $\Rd$ by $\Sd$. We write
$$ \ma \as x := \frac{\ma x}{\abs{\ma x}}, \qquad x \in \Sd$$ for the action of a matrix $\ma$ on $\Sd$ (as soon as this is well defined). If $\Sd$ is invariant under the action of $\mA_1$, we introduce a Markov chain on $\Sd$ by
$$ X_n^x := \mPi_n \as x, \qquad x \in \Sd.$$
\subsection{Nonnegative Matrices: Condition $\condC$}
Denote the cone of vectors with nonnegative entries by $\Rdnn$
and write
$$ \Sp = \{ x \in \Rdnn \, : \, \abs{x}=1\}$$
for its intersection with unit sphere.
The set of $d \times d$-matrices with nonnegative entries is denoted by $\Mset$ and we write
$$\interior{\Mset} = \{ \ma \in M(d \times d, \R) \, : \, \ma_{i,j}>0 \ \forall\, 1 \le i,j \le d \}$$ for its interior, which consists of matrices that have all entries positive.
A matrix $\ma \in \Mset$ is called \emph{allowable} (see \cite{Hennion1997}), if every row and every column has a positive entry.
If $\ma$ is an allowable matrix, then its action on $\Sp$ is well defined, and
moreover, the quantity
$$ \iota(\ma) := \min_{x \in \Sp} \abs{\ma x} > 0.$$
Consider now a probability distribution $\mu$ on $\Mset$. Write $[\supp \mu]$ for the subsemigroup generated by its support. We say that $\mu$ satisfies condition $\condC$, if:
\begin{enumerate}
\item Every $\ma \in [\supp \mu]$ is allowable.
\item $[\supp \mu] \cap \interior{\Mset} \neq \emptyset$.
\end{enumerate}
In the following, $\Gamma:= [\supp \mu]$. Observe that condition $\condC$ holds for $\Gamma$ if and only if it holds for $\Gamma^\top$.
Refering to the Perron-Frobenius theorem, every $\ma \in \interior{\Mset}$ possesses a unique dominant eigenvalue $\lambda_\ma$, (i.e.~$\abs{\lambda_\ma} > \abs{\lambda_i}$ for any other eigenvalue $\lambda_i$ of $\ma$) which is positive and algebraically simple, and a corresponding eigenvector $v_\ma \in \interior{\Sp}$. For a subsemigroup $\Gamma$ of allowable matrices, we define the collection of all such (normalized) dominant eigenvectors by
$$ V(\Gamma) := \closure{\left\{v_\ma \, : \, \ma \in \Gamma \cap \interior{\Mset}\right\}}. $$
It can be shown (see \cite[Lemma 4.3]{BDGM2014}) that $V(\Gamma)$ is the unique minimal $\Gamma$-invariant subset of $\Sp$, i.e.~every {\color{black} closed} $\Gamma$-invariant subset of $\Sp$ contains $V(\Gamma)$. It is worth mentioning already now, that the Markov chain $X_n^x$ possesses a unique stationary probability measure, the support of which is given by $V(\Gamma)$.
\subsection{Invertible Matrices: Condition \textit{(i-p)}}
In order to highlight connections, we decided to use the same symbols for objects which play the same role in the context of invertible matrices as they did for nonnegative matrices. The condition \textit{(i-p)} (irreducible and proximal), described below, is due to Guivarc'h, Le Page and Raugi and was studied in detail in several articles by these authors, the most comprehensive one of which is \cite{Guivarch2012}.
Let now $\mu$ be a probability measure on the group $GL(d,\R)$ of invertible $d \times d$ matrices and $\Gamma$ be the closed semigroup of $GL(d,\R)$ generated by $\supp \, \mu$. A matrix $\ma$ with an algebraic simple dominant $\lambda_\ma$ is called \emph{proximal}. This replaces the notion of a matrix with strictly positive entries, which is always proximal by the Perron-Frobenius theorem. Then the measure $\mu$ is said to satisfy condition \textit{(i-p)}, if
\begin{enumerate}
\item There is no finite union $\mathcal{W}=\bigcup_{i=1}^n W_i$ of subspaces $0 \neq W_i \subsetneq \R^d$ which is $\Gamma$-invariant, i.e. $\Gamma \mathcal{W}=\mathcal{W}$. (\emph{irreducibility})
\item $\Gamma$ contains a proximal matrix. (\emph{proximality})
\end{enumerate}
We will consider invertible matrices acting on the projective space $\mathbb{P}^{d-1}$ which is obtained from $\Sd$ by identifying $x$ with $-x$, i.e.
$$ \mathbb{P}^{d-1}~\simeq~ \Sd/\pm .$$ Studying the action of the matrices on $\mathbb{P}^{d-1}$ rather than on $\Sd$ has several technical advantages, for example, the definition
$$ V(\Gamma) := \closure{\left\{v_\ma \in \mathbb{P}^{d-1} \, : \, \ma \in \Gamma \text{ is proximal }\right\}}, $$
becomes unambiguous. Note that the norm $\abs{\ma x}$ for $x \in \mathbb{P}^{d-1}$ is well defined, since it does not depend on the choice of a representant of $x$ in $\Sd$.
For the case of invertible matrices, we have that
$$ \iota(\ma)~:=~ \inf_{x \in \mathbb{P}^{d-1}} \abs{\ma x} ~=~ \norm{\ma^{-1}}^{-1}.$$
\subsection{Invertible Matrices: Condition \textit{(id)}}
The third set of assumptions, called \textit{(id)}~for irreducible and density, appears first at the end of Kesten's work \cite{Kesten1973} and was elaborated by Alsmeyer and Mentemeier in \cite{AM2010}. In fact, it can be shown to imply condition \textit{(i-p)}. Due to the stronger assumption that $\mu$ is absolutely continuous, it often allows for simpler proofs, this is why we include it as an extra set of assumptions.
Let $\mu$ be a probability measure on $GL(d,\R)$ and $(\mA_n)_{n \in \N}$ be an i.i.d.~sequence with law $\mu$ and write $\mPi_n:=\mA_n \cdots \mA_1$. Then $\mu$ is said to satisfy condition $\textit{(id)}$ if
\begin{enumerate}
\item for all open $U \subset \Sd$ and all $x \in \Sd$, there is $n \in \N$ such that $\P{\mPi_n \as x \in U} >0$, and
\item there are a matrix $\ma_0 \in GL(d,\R)$, $\delta, c >0$ and $n_0 \in \N$ such that
$$
\P{\mPi_{n_0} \in d\ma } ~\ge~ c \1[{B_\delta(\ma_0)}](\ma) \, \llam(d \ma) ,$$
where $\llam$ denotes the Lebesgue measure on $\R^{d^2} \simeq M(d \times d,\R)$.
\end{enumerate}
The classical example is $\mu$ having a density about the identity matrix.
It is shown in \cite[Lemma 5.5]{AM2010} that $X_n^x$ is a Doeblin chain under condition \textit{(id)}. The support of its stationary probability measure is $\Sd$ by {\color{black} \cite[Proposition 4.3]{BDMM2013}}, therefore we are led to identify $V(\Gamma):=\Sd$ in the case of \textit{(id)}.
\subsection{Basic properties for all cases}\label{sect:defQ} Below, we identify $\S=\Sp$ in the case of nonnegative matrices, $\S=\mathbb{P}^{d-1}$ in the case of \textit{(i-p)}-matrices and $\S=\Sd$ in the case of \textit{(id)}-matrices. Given a measure $\mu$ on matrices, set
$$ I_\mu := \{ s \ge 0 \, : \, \int_{} \norm{\ma}^s \, \mu(\d \ma) <\infty \}.$$
Then, for $s \in I_\mu$, we define an operator in the set $\Cf{\S}$ of continuous functions on $\S$ by
\begin{equation}\label{eq:Ps} \Ps f(x) := \int_{} \abs{\ma x}^s \, f(\ma \as x) \, \mu(\d \ma), \end{equation}
and the 'transposed' operator by
\begin{equation}\label{eq:Pst} \Pst f(x) := \int_{} \abs{\ma^\top x}^s \, f(\ma^\top \as x) \, \mu(\d \ma). \end{equation}
Properties of both operators, which will be given in a moment, will be important in our results.
Beforehand, we introduce a function that will turn out to describe the spectral radius of these operators.
On $I_\mu$, define the log-convex function
\begin{equation}\label{def:ks}
k(s) ~:=~ \lim_{n \to \infty} \left(\E{\norm{\mA_n \ldots \mA_1 }^s} \right)^\frac{1}{n} {\color{black} ~=~ \lim_{n \to \infty} \left(\E{\norm{\mA_n^\top \ldots \mA_1^\top }^s} \right)^\frac{1}{n}}.
\end{equation} {\color{black} Here the second identity holds since $\norm{\ma}=\norm{\ma^\top}$ and the $(\mA_i)_{i\in \N}$ are i.i.d. }
We have the following result:
\begin{prop}\label{prop:transferoperators}
Assume that $\mu$ satisfies $\condC$, \textit{(i-p)}~or \textit{(id)}~ and let $s \in I_\mu$.
\begin{enumerate}
\item Then the spectral radii $\rho(\Ps)$ and $\rho(\Pst)$ both equal $k(s)$. \label{c1}
\item There is a unique normalized function $\es \in \Cf{\S}$ and a unique probability measure $\nus \in \Pset(\S)$ satisfying
$$ \Ps \es = k(s) \es \quad \text{ and } \quad \Ps \nus = k(s) \nus.$$
\item The function $\es$ is strictly positive and $\bar{s}:=\min\{s,1\}$-H\"older continuous and $\supp\, \nus = V(\Gamma)$. \label{suppnus}
\item {\color{black} If $\nust$ is a probability measure satisfying $\Pst \nust = k(s) \nust$, then there is $c>0$ such that \label{nust}
$$ \es(x) ~=~ c \int_\S \abs{\skalar{x,y}}^s \nust(dy).$$}
\item The function $s \mapsto k(s)$ is log-convex on $I_\mu$, hence continuous on $\interior{I_\mu}$ with left- and right derivatives. \label{c4}
\item The function $s \mapsto k(s)$ is analytic on $\interior{I_\mu}$. \label{c5}
\end{enumerate}
\end{prop}
{\color{black}
\begin{proof}[Source:]
Claims \eqref{c1}--\eqref{c4} were proved in \cite[Proposition 3.1]{BDGM2014} for nonnegative matrices, in \cite[Theorem 2.6 and Theorem 2.17]{Guivarch2012} for invertible matrices under condition \textit{(i-p)}~ and in \cite[Theorem 17.1]{Mentemeier2013a} under condition \textit{(id)}. The analycity of $k(s)$ in assertion \eqref{c5} is proved by using perturbation theory, and was proved first under condition \textit{(i-p)}~ in \cite[Corollary 3.20]{Guivarch2012}, and subsequently, using the same methods, in \cite[Corollary 4.12]{BDMM2013} under condition \textit{(id)}~ and is proved below in Corollary \ref{cor:k} for nonnegative matrices.
\end{proof}
\begin{rem}
Given only finiteness of $$\E \bigg( \big( 1+\norm{\mA}^{s_0} \big) \big(\abs{\log \norm{\mA}} + \abs{\log \iota(\mA)} \big) \bigg),$$ the mapping $s \mapsto k(s)$ is still differentiable on the {\em closed} interval $[0,s_0]$, this has been proved in \cite[Theorem 3.10]{Guivarch2012} under condition \textit{(i-p)}~ and in \cite[Theorem 6.1]{BDGM2014} for nonnegative matrices.
\end{rem}
}
Proposition \ref{prop:transferoperators} is crucial in order to define an exponential change of the measure $\mu$: Let $\Omega=M(d \times d, \R)^\N$ and $(\mA_n)_{n \in \N} : \Omega \to \Omega$ the fibered identity.
Now introducing for each $n$ the kernel
\begin{equation}
\label{qn}
q_n^s(x,\ma) = \frac{|\ma x|^s}{k^n(s)}\frac{\es(\ma\cdot x)}{\es(x)},
\end{equation}
we see that for each $x \in \S$ and $n \in \N$,
$$ \int q_n^s(x,\ma_n \cdots \ma_1) \mu^{\otimes n}(\d \ma_1, \dots, \d \ma_n)=1$$
{\color{black} and the relation
\begin{equation}\label{eq:qns}
q_{n}^s(x, \ma) q_{m}^s(\ma \as x, \mb) = q_{n+m}^s(x, \mb \ma).
\end{equation}
}
Moreover, for each $x \in \S$ the sequence $q_n^s(x, \cdot) \mu^{\otimes n}$ of probability measures is projective, hence by the Kolmogorov extension theorem, it gives rise to a probability measure ${\QQ_x^s}$ on $\Omega$, which we call the $s$-shifted measure. The corresponding expectation symbol is denoted by $\E_{{\QQ_x^s}}$. Note that $(\mA_n)_{n \in \N}$ are i.i.d.~with law $\mu$ for $s=0$. {\color{black} We use the symbol $\Q_x$ for $\Q_x^0$.}
With the conventions ${\QQ_x^s}(\{X_0 = x\})=1$, we have the Markov chain $X_n$ and the Markov additive process $S_n$:
\begin{eqnarray*}
X_n & :=& \mA_n \as X_{n-1} = \frac{\mA_n X_{n-1}}{\abs{\mA_n X_{n-1}}},\\
\quad S_n &:=& \log \abs{\mA_n \cdots \mA_1 X_0} {\color{black}~=~ \log \abs{\mA_n X_{n-1}} + S_{n-1} }.
\end{eqnarray*}
{\color{black} The second identity shows that $(X_n, S_n)$ carries the structure of a {\em Markov Random Walk}, i.e. the law of the increments $S_n - S_{n-1}$ depends on the past only via $X_{n-1}$. }
Writing as before $\mPi_n := \mA_n \cdots \mA_1$, we have the following fundamental identities, valid for any bounded measurable function $f$ and $n \in \N$:
\begin{align}
\frac{1}{k(s)^n \es(x)}\Erw{f(x,\mA_1, \dots, \mA_n) \es(X_n^x) \abs{\mPi_n x}^s} = & \Erw[{\QQ_x^s}]{f(X_0, \mA_1, \dots, \mA_n)}, \\
\frac{1}{k(s)^n \es(x)}\Erw{f\Bigl((X_k^x, S_k^x)_{k=0}^n \Bigr) \es(X_n^x) \abs{\mPi_n x}^s} = & \Erw[{\QQ_x^s}]{f\Bigl((X_k, S_k)_{k=0}^n \Bigr)}. \label{eq:com1}
\end{align}
The transition operator of $(X_n)_{n \in \N}$ is given by
\begin{equation}\label{eq:defQs} \Qs f(x) := \frac{1}{\es(x) k(s)} \Ps (f \cdot \es)(x). \end{equation}
It follows from Proposition \ref{prop:transferoperators} that $\Qs$ has a unique stationary probability measure
$$\pi^s := \frac{\es\nu^s}{\nu^s(\es)}$$ with support $V(\Gamma)$.
We set $$\QQ^s := \int {\QQ_x^s} \, \pis(dx).$$
\subsection{On $S_n^x$.}
Each of the assumptions introduced above is sufficient for the announced extension of the Furstenberg-Kesten theorem to hold:
\begin{prop}\label{prop:FK}
Assume that $\mu$ satisfies $\condC$ or \textit{(i-p)}~ or \textit{(id)}. Let $s \in \{0\} \cup \interior{I_\mu}$ and assume there is $0 < \epsilon < 1$ such that
\begin{equation}\label{eq:iotamoment}
\E \norm{\mA}^{s+\epsilon} \iota(\mA)^{-\epsilon} < \infty.
\end{equation}
Then it holds that $q:= \E_{\QQ^s} S_1 = k'(s)/k(s) \in \R$, and
$$ {\color{black} \lim_{n \to \infty} \frac{1}{n} \log \norm{\mPi_n} ~=~} \lim_{n \to \infty} \frac{S_n}{n} ~=~ q \qquad {\QQ_x^s}\text{-a.s.}$$
for all $x \in \S$.
\end{prop}
This is proved in \cite[Theorem 6.1]{BDGM2014} under condition $\condC$, in \cite[Theorem 3.10]{Guivarch2012} under condition \textit{(i-p)}~ and in \cite[Proposition 20.2]{Mentemeier2013a} under condition \textit{(id)}. In the last reference, the first identity is not proved, but it follows from the corresponding result for \textit{(i-p)}.
\begin{rem}
Recall from Proposition \ref{prop:transferoperators} that $k(s)$ is log-convex. Therefore,
$$ \Lambda(s)~:=~ \log k(s)$$ is convex, and
$$ q ~=~ \frac{k'(s)}{k(s)} ~=~ \Lambda' (s) ~\ge~ \Lambda'(0) ~=~ \frac{k'(0)}{k(0)} ~=~\gamma. $$
The function
$$ \Lambda^*(q) ~:=~ sq - \Lambda(s) $$
is the Fenchel-Legendre transform of $\Lambda$ and nondecreasing on $I_\mu$, see \cite[Lemma 2.2.5]{Dembo1998}. In particular, it is nonnegative on $I_\mu$.
\end{rem}
When studying random walks, an important distinction is between so-called lattice types, i.e. whether or not the random walk takes values only in some lattice $c \Z$ for $c \ge 0$ . A similar concept applies for Markov random walks, which are introduced below. The lattice type of $S_n$ only depends on the support of $\mu$, thus we give first a measure-free definition, which implies the more frequently used subsequent definition, which is relative to the measure {\color{black} $\Q^s$}.
\begin{defin}\label{lem:non-arithmetic}
\begin{enumerate}
\item We say that $\Gamma$ resp. $\mu$ is \emph{arithmetic}, if there is $t>0$ together with $\theta \in [0, 2\pi)$ and a function $\vartheta : \Sp \to \R$ such that
\begin{equation}\label{eq:non-arithmetic}\tag{A}
\forall \ma \in \Gamma, \ \forall x \in V(\Gamma) \ : \ \exp\Bigl(i t \log \abs{\ma x} - i \theta + i (\vartheta(\ma \as x) - \vartheta(x)) \Bigr) =1.
\end{equation}
If no such $t$ exists, then $\Gamma$ is said to be \emph{non-arithmetic}.
\item The Markov random walk $(X_n, S_n)$ is said to be arithmetic under {\color{black} $\Q^s$}, if there is $t>0$ together with $\theta \in [0, 2\pi)$ and a function $\vartheta : \S \to \R$ such that
\begin{equation}\label{eq:arithmetic2
\E_{\Q^s}\exp\Bigl(i t S_1 - i \theta + i (\vartheta(X_1) - \vartheta(X_0)) \Bigr) =1,
\end{equation}
and non-arithmetic otherwise.
\end{enumerate}
\end{defin}
We have the following implications.
\begin{lem}\label{lem:implications_arithmetic}
If $\Gamma ~=~ [\supp \mu]$ is arithmetic, then $(X_n, S_n)$ is arithmetic under each {\color{black} $\Q^s$} with the same $t, \theta, \vartheta$.
Conversely, if $(X_n, S_n)$ is arithmetic under some {\color{black} $\Q^s$} and the function $\vartheta$ is continuous on $\S$, then $\Gamma$ is arithmetic as well with the same $t, \theta, \vartheta$.
\end{lem}
\begin{proof}
Recalling that $\supp \pis = V(\Gamma)$, we observe that Eq. \eqref{eq:arithmetic2} is equivalent to
$$ \exp\Bigl(i t \log{\ma x} - i \theta + i (\vartheta(\ma \as x) - \vartheta(x)) \Bigr) ~=~1 \qquad \text{ for $\mu$-a.e. $\ma \in \supp\, \mu$ and $\pis$-a.e. $x \in V(\Gamma)$ },$$
i.e. for dense subsets of $\supp\, \mu$ resp. $V(\Gamma)$, which gives the asserted implications.
\end{proof}
It is shown in \cite[Proposition 4.6]{Guivarch+Urban:2005} that under condition \textit{(i-p)}, $\Gamma=[\supp\, \mu]$ is non-arithmetic, while it is shown in \cite[Lemma 5.8]{AM2010}, that $(X_n, S_n)$ is {\color{black} non-arithmetic} under each {\color{black} $\Q^s$} under condition $\textit{(id)}$.
A simple sufficient condition (due to Kesten \cite{Kesten1973}) for $\Gamma$ to be non-arithmetic under condition $\condC$ is the following.
Set $$ S(\Gamma) := \{\log \lambda_\ma \ : \ \ma \in \Gamma \cap \interior{\Mset} \}.$$
\begin{lem}
Assume that the (additive) subgroup of $\R$ generated by $S(\Gamma)$ is dense.
Then $\mu$ is non-arithmetic.
\end{lem}
\begin{proof}
Supposing that Eq. \eqref{eq:non-arithmetic} holds for some $t, \theta$ and $\vartheta$, then we have for any $\ma \in \Gamma \cap \interior{\Mset}$ that $v_\ma \in V(\Gamma)$, hence
$$ \exp\Bigl(i\bigl[ t \log \abs{\ma v_\ma} - \theta +(\vartheta(\ma \as v_\ma) - \vartheta(v_\ma)) \bigr] \Bigr] =
e^{i(t\log \lambda_{\ma} - \theta)}
$$
Consequently, for any $\ma, \matrix{h} \in \Gamma \cap \interior{\Mset}$,
$$\log \lambda_\ma -\log \lambda_\matrix{h} \in \frac{2 \pi}{t} \Z.$$
But by our assumption, $S(\Gamma)$
is not contained in $\frac{2 \pi}{t} \Z$ for any $t >0$; this gives a contradiction.
\end{proof}
\begin{cor}
If there are $\ma, \mb \in \Gamma \cap \interior{\Mset}$ with $\frac{\log \lambda_\ma}{\log \lambda_\mb} \notin \Q$, then $\mu$ is non-arithmetic.
\end{cor}
Now we have enough notation to state our main results.
\section{Statement of main results}
We will prove the following analogue of the Bahadur-Rao theorem for products of random matrices. The role of the cumulant generating function is played here by $\Lambda(s) =\log k(s)$.
\begin{thm}\label{thm:BahadurRao}
Assume that $\mu$ satisfies $\condC$ and is non-arithmetic, or that $\mu$ satisfies \textit{(i-p)}~ or \textit{(id)}.
If $q=\E_{\QQ^s} S_1 = \Lambda'(s) $ for some $s \in \interior{I_\mu}$ and there is $0 < \epsilon < 1$ such that
\eqref{eq:iotamoment} holds, then
$$ \lim_{n \to \infty}\, \sup_{x \in \S} \abs{ \sqrt{n}\, e^{n\Lambda^*(q)}\, J(s)\, \Erw{\es(X_n^x) \1[\{ S_n^x \ge n q \}]} - \es(x)} ~=~0, $$
where $$ J(s) ~=~ {s \sigma \sqrt{2 \pi} }, \qquad \text{ with } \sigma^2 ~=~ \Lambda''(s) ~=~ \lim_{n \to \infty} \frac{1}{n} \E_{\QQ^s} (S_n-nq)^2 > 0.$$\end{thm}
Since the function $\es$ is strictly positive and continuous on the compact set $\S$, hence bounded, this gives in particular uniform bounds for the large deviation probabilities:
\begin{cor}
There are $0 < c \le C < \infty$ such that for all $x \in \S$,
$$ c ~\le~ \liminf_{n \to \infty}~ \sqrt{n}\,(e^{sq})^n\,\P{ S_n^x \ge n q } ~\le~ \limsup_{n \to \infty}~ \sqrt{n}\,(e^{sq})^n\,\P{ S_n^x \ge n q } ~\le~ C. $$
\end{cor}
These large deviations results will be used to prove the following result about random difference equations, which gives an elementary proof that the tail estimates derived e.g. in \cite{Kesten1973,Klueppelberg2004,AM2010,Guivarch2012} are precise:
\begin{thm}\label{thm:rde}
Let {\color{black} $\mM$ be a random matrix and let $B$ be a random vector in $\Rd$. Write $\mA:=\mM^\top$ and denote by $\mu$ the law of $\mA$. Assume that $k'(0)<0$ and that there is $\alpha \in \interior{I_\mu}$ with $k(\alpha)=1$ and }
\begin{equation}\label{eq:moment conditions}
\E \norm{\mA}^{\alpha+\epsilon} \iota(\mA)^{-\epsilon} < \infty, \qquad 0 < \E \abs{B}^{\alpha + \epsilon} < \infty
\end{equation}
for some $\epsilon >0$. There is a random variable $R$, unique in distribution, satisfying {\color{black} $R \eqdist \mM R + B$}.
\begin{enumerate}
\item \label{nonnegative} Let {\color{black} $\mA$} be nonnegative, satisfying condition $\condC$ and being non-arithmetic. Assume that $\supp R \cap \Rdnn$ is unbounded. Then there is $\delta > 0$ such that for all $x \in \Sp$,
$$ \liminf_{t \to \infty} \, t^\alpha \P{\skalar{x,R}>t} \ge \delta.$$
\item Let $\mA \in GL(d,\R)$, satisfying \textit{(id)}. Assume that $\P{\mA r + B = r} < 1$ for all $r \in \Rd$.
Then there is $\delta > 0$ such that for all $x \in \Sd$,
$$ \liminf_{t \to \infty} \, t^\alpha \P{\skalar{x,R}>t} \ge \delta.$$
\item {\color{black} Let $\mA \in GL(d,\R)$, satisfying \textit{(i-p)}. Assume that $\Gamma^*$ does not leave invariant any proper closed convex cone in $\Rd$, and that $\P{\mA r + B = r} < 1$ for all $r \in \Rd$.
Then there is $\delta > 0$ such that for all $x \in \Sd$,
$$ \liminf_{t \to \infty} \, t^\alpha \P{\skalar{x,R}>t} \ge \delta.$$}
\end{enumerate}
\end{thm}
In this theorems, {\color{black} we impose the assumptions on the law of $\mA=\mM^\top$ rather than on the law of $\mM$ (note nevertheless, that $\condC$ or \textit{(i-p)}~ hold for $\mM^\top$ as soon as they hold for $\mM$). The reason is as follows: Let $(\mM_k, B_k)_{k \in \N}$ be a sequence of i.i.d.~copies of $(\mM,B)$. Then, upon iterating Eq. \eqref{eq:rde1}, we obtain {\color{black} $R \eqdist \mM_1 \cdots \mM_n R + \sum_{k\le n} \mM_1 \cdots \mM_{k-1}B_k$}, which leads to the study of
$$ \skalar{x,R} ~\eqdist~ \skalar{x,\mM_1 \cdots \mM_n R + \sum_{k\le n} \mM_1 \cdots \mM_{k-1}B_k} ~=~ \skalar{\mM_n^\top \cdots \mM_1^\top x, R} + \ldots,$$
and we are going to show that the first term dominates in order to use Theorem \ref{thm:BahadurRao} to derive estimates.}
\begin{rem}\label{rem:rde}
Let us stress that in \eqref{nonnegative} we do not assume that $B$ is nonnegative and that the condition $\supp R \cap \Rdnn$ being unbounded is obviously also necessary for the heavy tail property. Thereby, we generalize the result of Kesten, namely \cite[Theorem 3]{Kesten1973}.
A sufficient condition for $\supp R \cap \Rdnn$ being unbounded is $B$ being nonnegative, or {\color{black} $\mM,B$} being independent and $\P{B \in \Rdp} > 0$.
\end{rem}
\begin{rem
The law of the random variable $R$ is given by
{\color{black} $ \sum_{k=1}^\infty \mM_1 \cdots \mM_{k-1} B_k,$}
from which we immediately obtain the estimate (for $s \ge 1$)
$$ {\color{black} (\E \abs{R}^s)^{1/s} ~ \le ~ \sum_{k=1}^\infty \left( \E \norm{\mM_1 \cdots \mM_n}^s \right)^{1/s} \, (\E \abs{B}^s)^{1/s} ~=~ \sum_{k=1}^\infty \left( \E \norm{\mM_1^\top \cdots \mM_n^\top}^s \right)^{1/s} \, (\E \abs{B}^s)^{1/s}.}$$
This shows that if {\color{black} $k(s) <1$} and $\E \abs{B}^s < \infty$, then readily $\E \abs{R}^s < \infty$, which shows in particular that under the assumptions of Theorem \ref{thm:rde},
$$ \limsup_{t \to \infty} t^{s} \P{\skalar{x,R}>t} = 0 $$
for all $0 \le s < \alpha$ and all $x \in \mathcal{S}$.
\end{rem}
{\color{black}
\begin{rem}
The moment conditions \eqref{eq:moment conditions} are not optimal, precise tail estimates have been obtained
under the assumptions
$$ \E \norm{\mA}^\alpha ( \log \norm{\mA} + \abs{ \log \iota(\mA)}) < \infty, \qquad 0 < \E \abs{B}^\alpha < \infty, $$
see \cite[Remark after Theorem 5.2]{Guivarch2012} in the case of \textit{(i-p)}~resp.~ \cite[Theorem 13.2]{Mentemeier2013a} for the case of condition \textit{(id)}.
\end{rem}
}
\subsection{Structure of the paper and sketch of proofs}
The proof of Theorem \ref{thm:BahadurRao}
will rest upon a third-order Edgeworth expansion for the cdf
$$ F_{n,x}^s(t) := {\QQ_x^s}\left\{ \frac{S_n - nq}{\sigma \sqrt{n}} \le t \right\},$$ which is given in Theorem \ref{thm:edgeworth}.
To prove this intermediate result, we will use the Nagaev-Guivarc'h spectral method as in Hennion and Herv\'e \cite{HH2001} and Herv\'e and Pen\`e \cite{Herve2010}: The classical Edgeworth expansion for random walks
can be proved using the Fourier transform of $S_n$, in particular its behavior at zero.
{\color{black} Upon introducing (for suitable $z \in \C$) the operator $Q(z)$ in $\Cf{\S}$ by
$$Q(z) f(x) ~:=~ \frac{1}{\es(x) k(s)} \int_{\Gamma} \abs{\ma x}^{s +z} f(\ma \as x) \es(\ma \as x) \, \mu(\d \ma)
$$
we have the following fundamental identity for the Fourier transform $\phi_{n,x}$ of ${\QQ_x^s}\{S_n \in \cdot \}$:
\begin{equation} \phi_{n,x}(t):= \E_{\Q_x^s}\left( e^{itS_n}\right) = \E_{\Q_x^s}\left( e^{itS_n} \, \1[\S](X_n) \right) = Q(it)^n \1[\S](x). \label{eq:FT} \end{equation}
This identity is a consequence of the following lemma.
\begin{lem}\label{lem:FT}
Let $\mA$ be a random matrix with law $\mu$, and assume that $\E \norm{\mA}^{s+\Re z} < \infty$ for $s >0, z \in \C$. Then the following identity holds for all $f \in \Cf{\S}$:
\begin{equation} \E_{\Q_x^s}\left( e^{zS_n} f(X_n)\right) ~=~ Q(z)^n f(x) \label{eq:Qzt} \end{equation}
\end{lem}
\begin{proof} The assumption guarantees that $Q(z)$ is well defined, and all integrals appearing below are finite.
We use induction. For $n=1$, this is immediate from the definition of $Q(z)$ and identity \eqref{eq:com1}. Suppose \eqref{eq:FT} holds for $n \in \N$. Then, using again \eqref{eq:com1} and the fact that the $(\mA_i)$ are i.i.d~ with law $\mu$ under $\Prob$, we obtain
\begin{align*}
Q(z)^{n+1}f(x) ~&=~ Q(z) \left( Q(z)^n f)(x) \right) \\
&=~ \int_{\Gamma} \frac{\es(\ma \as x) \abs{\ma x}^{s+z}}{\es(x) k(s)} \E_{\Q_{\ma \as x}^s} (e^{z S_n} f(X_n)) \, \mu(\d \ma) \\
&=~ \int_{\Gamma} \frac{\es(\ma \as x) \abs{\ma x}^{s+z}}{\es(x) k(s)} \frac{1}{\es(\ma \as x) k(s)^n} \E \left( \abs{\mPi_n (\ma \as x)}^{s+z} \es( \mPi_n \as (\ma \as x)) f( \mPi_n \as (\ma \as x)) \right) \, \mu(\d \ma) \\
&=~ \frac{1}{\es(x) k(s)^{n+1}} \int_{\Gamma} \E \left( \abs{\mPi_n \ma x}^{s+z} \es( \mPi_n\ma \as x) f( \mPi_n\ma \as x) \right) \, \mu(\d \ma) \\
&=~ \frac{1}{\es(x) k(s)^{n+1}} \E \left( \abs{\mPi_{n+1} x}^{s+z} \es( \mPi_{n+1} \as x) .\, f( \mPi_{n+1} \as x) \right) \, \mu(\d \ma) \\
&=~ \E_{\Q_x^s} \left( e^{z S_{n+1}} f(X_{n+1})\right).
\end{align*}
\end{proof}
Observe that $Q(0)=Q^s$ and that, given $s \in \interior{I_\mu}$, the mapping $z \mapsto Q(z)$ is holomorphic in some domain. }We are going to show that the operator $\Qs$ is quasi-compact with a simple dominant eigenvalue $\theta(0)=1$, and thereupon, using holomorphic perturbation theory, the decomposition
$$ Q^n(z) = \theta(z)^n M(z) + L(z)^n,$$
for a rank-one projection $M$ and an operator $L(z)$ with spectral radius $\rho(L(z)) < \rho(Q(z))$.
From this we will finally deduce that for $n \to \infty$,
$$ \phi_{n,x}(t/\sqrt{n}) = Q^n(it/\sqrt{n}) \1[\S](x) \approx \theta(it\sqrt{n}),$$
i.e. behavior at zero of the Fourier transforms is given by small perturbations of the dominant eigenvalue of $\Qs$.
\medskip
Therefore, we start our investigations by proving spectral properties of $\Qs$ and the family $Q(z)$ (in the case of nonnegative matrices). In Section \ref{sect:quasi-compact}, we prove, continuing \cite{BDGM2014} and based on the approach in \cite{Guivarch2012}, that $Q^s$ is quasi-compact. This property is needed in order to apply a perturbation theorem which proves the decomposition of the family $Q(z)$ in Section \ref{sect:perturbation}. Then we are ready to prove a third-order Edgeworth expansion for $F_{n,x}^s$ in Section \ref{sect:edgeworth}, which is used to prove Theorem \ref{thm:BahadurRao} in Section \ref{sect:BahadurRao}. Sections \ref{sect:nonarithmetic} and \ref{sect:taylor} study the implications of the non-arithmeticity condition, as well as formulas for $\sigma^2$.
Section \ref{sect:rde} is concerned with Theorem \ref{thm:rde}. We start by providing an example, namely the ARCH(q)-process, to which our results apply and continue by giving an outline of the proof of Theorem \ref{thm:rde}, while we postpone the technical details to the final Section \ref{sect:rde proofs}.
\section{Quasi-compactness of $\Qs$}\label{sect:quasi-compact}
\subsection{Nonnegative matrices} In this section, which is based on the approach of Guivarc'h and Le Page \cite{Guivarch2012} for \textit{(i-p)}, we are going to prove that for each $s \in I_\mu$, the operator $\Qs$ is quasi-compact (has a spectral gap) on a subspace of $\Cf{\Sp}$, namely the space of functions that are $\bar{s}:=\min\{s,1\}$-H\"older continuous with respect to a particular metric $d$ on $\Sp$. At first, we will recall the Theorem of Ionescu Tulcea and Marinescu, which will be used in order to prove the quasi-compactness. Then we introduce the particular metric $d$ which will be useful when finally checking the assumptions of this theorem.
\medskip
We write $\mathcal{L}(\B,\B)$ for the set of all bounded linear operators from $\B$ to $\B$. An operator $Q \in \mathcal{L}(\B,\B)$ is said to be {\em quasi-compact} if $\B$ can be decomposed into two {\color{black} closed} $Q$-invariant subspaces $\B = E \oplus F$ where the spectral radius $\rho(Q_{|F}) < \rho(Q)$ while $\dim E < \infty$ and each eigenvalue of $Q_{|E}$ has modulus $\rho(Q)$.
Subsequently, a convenient way to prove the quasi-compactness of $\Qs$ will be to use the following generalization of the {\color{black} Theorem of Ionescu-Tulcea and Marinescu}:
\begin{thm}[{\cite[Theorem II.5]{HH2001}}]\label{thm:ITM}
Let $(\B, \norma{\cdot})$ be a Banach space {\color{black} and let $\normb{\cdot}$ be a continuous semi-norm on $\B$. Assume that $Q$ is a bounded} operator in $\B$ such that
\begin{enumerate}
\item $Q\, \{f \, : \, \norma{f} \le 1 \}$ is conditionally compact in $(\B, \normb{\cdot})$. \label{prop1}
\item there exists a constant $M$ such that for all $f \in \B$, $\normb{Q f} \le M \normb{f}$, \label{prop2}
\item there exist $k \in \N$ and real numbers $r$ and $R$ with $r < \rho(Q)$ and, for all $f \in \B$,\label{prop3}
$$ \norma{Q^k f} \le R \normb{f} + r^k \norma{f}.$$
\end{enumerate}
Then $Q$ is quasi-compact.
\end{thm}
Though we have not yet defined the metric $d$ on $\Sp$, let us nevertheless state right now, which Banach space and what norms we are going to consider. For $f \in \Cf{\Sp}$, set
$$ \normb{f}:= \sup_{x \in \Sp} \abs{f(x)}, \qquad \abs{f}_s :=\sup_{x,y \in \Sp} \frac{\abs{f(x) - f(y)}}{d(x,y)^{\bar{s}}}, \qquad \norma{f}:=\normb{f}+\abs{f}_s.$$
We consider the Banach space
$$ \B := \{ f \in \Cf{\Sp} \, : \, \abs{f}_s < \infty \} = \{ f \in \Cf{\Sp} \, : \, \norma{f} < \infty \}$$
equipped with the norm $\norma{\cdot}$.
Using
Theorem \ref{thm:ITM}, we are going to prove the following:
\begin{prop}\label{prop:Qs}
Assume that $\mu$ {\color{black} satisfies} $\condC$ and let $s \in I_\mu$. Then $Q^s \in \mathcal{L}(\B,\B)$, and there is an operator $N \in \mathcal{L}(\B,\B)$ with spectral radius $\rho(N) < 1$, such that
\begin{equation} \label{eq:decompQs} (Q^s)^n = M + N^n \end{equation}
for all $n \in \N$, where $M$ is a rank-one projection onto $\R\1[\Sp]$ with $M(f)(x) = \pis(f)$ for all $f \in \B$ and $x \in \Sp$.
\end{prop}
This will be done by a series of Lemmata, which will make use of the particular metric $d$ on $\Sp$, which we are going to introduce next.
\subsubsection{A metric on $\Sp$}\label{subsect:metric}
Given $x \neq y \in \Sp$, consider the line $L$ trough these points. {\color{black} Then $L \cap \partial \Rdnn$ consists of two points which we label by $a$ and $b$ in such a way that if we write $x= u_1 a + u_2 b$ and $y=v_1 a + v_2 b$ $u_1,u_2,v_1,v_2 \ge 0$ as convex combinations of $a$ and $b$, then $u_1 > v_1$, i.e. $x$ lies between $a$ and $y$. Then the cross-ratio of $a,b$ and $x,y$ is given as
$$ [a,b;x,y] = \frac{u_2 v_1}{u_1 v_2} .$$}
The formulae $$d(x,y):=\phi([a,b;x,y])$$ for $\phi(s):= \frac{1-s}{1+s}$, $s \in [0,1]$,
defines a bounded distance on the unit sphere.
Its properties are summarized in the following {\color{black} Proposition}.
\begin{prop} \label{prop:properties of d}
For any norm $\abs{\cdot}$, $d$ is a metric on $\Sp$ with
\begin{itemize}
\item $\sup\{d(x,y) \, : \, x,y \in \Sp \} =1$,
\item There is $C>0$ s.t. $d(x,y) \ge C \abs{x-y}$.
\end{itemize}
For $\ma \in \Mset$, there exists $c(\ma) \le 1$ such that:
\begin{enumerate}
\item $d(\ma \as x, \ma \as y) \le c(\ma) d(x,y),$
\item $c(\ma) <1$ if and only if $\ma \in \interior{\Mset}$,
\item if $\ma' \in \Mset$, then $c(\ma \ma') \le c(\ma) c(\ma')$,
\item $c(\ma^\top)=c(\ma)$.
\end{enumerate}
\end{prop}
{\color{black} \begin{proof}[Source:] This is \cite[Proposition 3.1]{Hennion1997}.
There, the results are stated relative to the 1-norm $\norm{x}_1 = \sum_{i=1}^n \abs{x_i}$ on $\Rd$, but they do in fact hold for any norm on $\Rd$, the main reason being that the cross-ratio is an projective invariant and thus independent of the shape of the unit sphere, and that all norms on $\Rd$ are comparable.
\end{proof}
}
The crucial properties of the metric $d$ are (1) and (2), saying that the action of nonnegative (positive) matrices is a (strict) contraction with respect to $d$.
\subsubsection{Checking the assumptions of the Ionescu-Tulcea-Marinescu theorem}
Let us first recall the definition of $\Qs$ in \eqref{eq:defQs}, from which we obtain the following formula for its iterates:
\begin{equation}
\label{eq:Qsn} (\Qs)^n f(x) ~=~ \Erw[{\QQ_x^s}]{f(X_n)} ~=~ \Erw{q_n^s(x,\mPi_n) f(\mPi_n \as x)}.
\end{equation}
In order to prove Assumption (1) of Theorem \ref{thm:ITM}, we are going to apply the Arzel\`a-Ascoli theorem. Therefore, we have to prove equicontinuity of the family $\{ \Qs f \, : \, \norma{f} < \infty \}.$
This will follow from the subsequent estimates for the kernels $q_n^s$, {\color{black} where it is shown} in particular, that the mappings $q_n^s(\cdot, \ma)$ are $\bar{s}$-H\"older on $\Sp$ for any {\color{black} $\ma \in \Mset$}.
\begin{lem}\label{lem:prop:qn}
{\color{black} Under the assumptions of Proposition \ref{prop:Qs}, there is} $C_s < \infty$ such that for all $n \in \N$, $x,y \in \Sp$, $\ma \in \Mset$
$$ \abs{q_n^s(x,\ma) - q_n^s(y,\ma)} \le C_s \frac{\norm{\ma}^s}{k(s)^n} d(x,y)^{\bar{s}}.$$
On the other hand, there is $c_s$ such that for all allowable $\ma$,
{\color{black} $$ q_n^s(\ma) := \int q_n^s(x, \ma) \pis(\d x) \ge \frac{c_s}{k(s)^n} \norm{\ma}^s$$}
\end{lem}
\begin{proof
Observe that by Proposition \ref{prop:properties of d}, any function that is H\"older-continuous on $(\Sp, \abs{\cdot})$ is as well H\"older-continuous on $(\Sp, d)$.
Using that thus $\es$ is $\bar{s}$-H\"older with constant $d_{\es}$ and bounded with $0<d_1 \le \es(x) \le d_2<\infty$ for all $x \in \Sp$, as well as property (2) of Proposition \ref{prop:properties of d}, we estimate
\begin{align*}
& \abs{\frac{\es(\ma \as x)}{\es(x)} \frac{\abs{\ma x}^s}{k(s)^n} - \frac{\es(\ma \as y)}{\es(y)} \frac{\abs{\ma y}^s}{k(s)^n}} \\
\le~& \abs{\frac1{\es(x)} - \frac{1}{\es(y)}} \frac{\es(\ma \as x) \abs{\ma x}^s}{k(s)^n} + \abs{\abs{\ma x}^s - \abs{\ma y}^s} \frac{\es(\ma \as x) }{\es(y) k(s)^n}
+ \abs{\es(\ma \as x) - \es(\ma \as y)} \frac{\abs{\ma y}^s}{k(s)^n \es(y)} \\
\le~& \frac{1}{d_1^2} \abs{\es(x)-\es(y)} \frac{d_2 \norm{\ma}^s}{k(s)^n} + \abs{\abs{\ma x}^s -\abs{\ma y}^s} \frac{d_2}{d_1 k(s)^n} + d_{\es} \abs{x-y}^{\bar{s}} \frac{\norm{\ma}^s}{k(s)^n d_1} \\
\le~& \left(\frac{C d_{\es} d_2}{d_1^2} + \frac{C d_{\es}}{d_1} \right) \frac{\norm{\ma}^s}{k(s)^n} d(x,y)^{\bar{s}} + \frac{d_2}{d_1 k(s)^n)} \abs{\abs{\ma x}^s - {\color{black} \abs{\ma y}^s}}.
\end{align*}
The last term has to be estimated differently for $s \le 1$ and $s >1$. If $s \le 1$, then
$$ \abs{\abs{\ma x}^s - {\color{black} \abs{\ma y}^s}} \le \abs{\abs{\ma x} - \abs{\ma y}}^s \le \norm{\ma}^s \abs{x -y}^s.$$
If $s >1$, then
$$ \abs{\abs{\ma x}^s - {\color{black} \abs{\ma y}^s}} \le \abs{\abs{\ma x} - \abs{\ma y}} \cdot s \cdot \max\{\abs{\ma x}^{s-1}, \abs{\ma y}^{s-1} \} \le s\norm{\ma} \abs{x -y}^{\bar{s}} \norm{\ma}^{s-1} $$
For the second part, recall $K = \inf\{ \es(x) / \es(y) \, : \, x,y \in \Sp \} >0$,
hence
\begin{align*}
q_n^s(\ma) \ge \frac{K}{k(s)^n} \int \abs{\ma x}^s \pis(\d x).
\end{align*}
It suffices to prove that $g(\ma):= \int \abs{\ma x}^s \pis(\d x) \ge c_s$ for all nonnegative $\ma$ with $\norm{\ma}=1$. On the compact set $\norm{\ma}=1$, $g$ attains its infimum. But if there is $\ma_0$ with $\int \abs{\ma_0 x}^s \pis(\d x) = 0$, then
$$ V(\Gamma) \subset \supp \nus = \supp \pis \subset \mathrm{ker}(\ma_0). $$
But since $\ma_0$ is a nonzero nonnegative matrix, $\mathrm{ker}(\ma_0) \cap \interior{\Sp} = \emptyset$, which gives a contradiction.
\end{proof}
{\color{black} Let us note the following, surprising Corollary to Lemma \ref{lem:prop:qn}, which shows that the convergence in \eqref{def:ks} is exponentially fast.
\begin{cor}\label{cor:ksEs}
Under the assumptions of Proposition \ref{prop:Qs}, there is $c_s >0$ (the same as in Lemma \ref{lem:prop:qn}), such that
$$ k(s)^n \le \E \norm{\mPi_n}^s \le \frac{1}{c_s} k(s)^n \qquad \text{ for all $n \in \N$}.$$
\end{cor}
\begin{proof}
The first inequality holds since $k(s) = \lim_{n \to \infty} \big(\E \norm{\mPi_n}^s \big)^{1/n} = \inf_{m \in \N} \big(\E \norm{\mPi_m}^s \big)^{1/m}$ due to submultiplicativity of the norm (see \cite[Theorem 1]{Furstenberg1960} for details). The second inequality holds by Lemma \ref{lem:prop:qn}, since $\E q_n^s(x,\mPi_n) =1$ for all $x \in \Sp$.
\end{proof}
}
Now we are ready to prove the following estimate, from which the validity of assumptions (1) and (3) will follow.
\begin{lem}\label{lem:hoeldercontinuous}
{\color{black} Under the assumptions of Proposition \ref{prop:Qs}, there is} $C >0$ and a sequence $D(n)$ with $\lim_{n \to \infty} D(n) = 0$, such that for all $n \in \N$ and $f \in \B$,
\begin{align}\label{eq:hoeldercontinuous}
\abs{(Q^s)^n f}_s \le C \normb{f} + D(n) \abs{f}_s
\end{align}
\end{lem}
\begin{proof}
For all $f \in \B$, $\abs{f}_s < \infty$. For such $f$, we compute
\begin{eqnarray*}
\abs{(\Qs)^n f(x) - (\Qs)^n f(y)}
&= & \abs{ \E_{{\QQ_x^s}} f(X_n) - \E_{{\QQ_y^s}} f(X_n)} \\
&\le& \E_{\QQ_x^s}\abs{ f(\mPi_n \as x) - f(\mPi_n \as y) } + \abs{(\E_{\QQ_x^s} - \E_{\QQ_y^s}) f(\mPi_n \as y)}\\ &=& I + II.
\end{eqnarray*}
Considering $I$,
\begin{align*}
I \le \abs{f}_s\, \E_{\QQ_x^s} d(\mPi_n \as x, \mPi_n \as y)^{\bar{s}} \le \abs{f}_s d(x,y)^{\bar{s}} \ \E_{\QQ_x^s} c(\mPi_n)^{\bar{s}}.
\end{align*}
But due to Proposition \ref{prop:properties of d}, $c(\ma) \le 1$ for all $a \in \Mset$, and $c(\ma) < 1$ for $\ma \in \interior{\Mset}$. {\color{black} By $\condC$, we have that
$\P{ \liminf_{n \to \infty } \left\{ \mPi_n \in \interior{\Mset} \right\}} = 1$ (see \cite[Lemma 3.1]{Hennion1997}), thus $c(\mPi_n) \to 0$ ${\QQ_x^s}$-a.s.~ by Proposition \ref{prop:properties of d}, (2) and (3) and the boundedness of $c$. Moreover, $c$ is continuous on $\Mset$ by \cite[Lemma 10.8]{Hennion1997}. Therefore, we can use the dominated convergence theorem to infer
$$ \lim_{n \to \infty} D(n) := \lim_{n \to \infty} \E_{\QQ_x^s} c(\mPi_n)^{\bar{s}} = 0.$$}
Turning to $II$, we have, using Lemma \ref{lem:prop:qn},
$$ II \le [f] \E \big| q_n^s(x,\mPi_n) - q_n^s(y,\mPi_n) \big|
\le \frac{[f] C_s d(x,y)^{\bar{s}}}{k(s)^n} \E \norm{\mPi_n}^s
\le [f] \frac{C_s}{c_s} d(x,y)^{\bar{s}} \E_{\QQ^s} \1
$$
Combining these estimates, we arrive at
\begin{align*}
\abs{(Q^s)^n f}_s \le \frac{C_s}{c_s} \normb{f} + \abs{f}_s
D(n).
\end{align*}
\end{proof}
\begin{proof}[Proof of Proposition \ref{prop:Qs}]
Now we are ready to show that Theorem \ref{thm:ITM} applies for $Q=\Qs$ with $\B$ as defined above.
\Step[1]:
Assumption \eqref{prop2} is satisfied for $M=1$, since $\Qs$ is a Markov operator on $(\Cf{\Sp},\normb{\cdot})$.
\medskip
\Step[2]: Assumption \eqref{prop1} holds for $\Qs$, i.e. $\Qs\{ f \, : \, \norma{f} \le 1 \}$ is conditionally compact in $(\B, \normb{\cdot})$. This is shown as follows. Since $\Qs$ is a Markov operator, and $\norma{f} \ge \normb{f}$, we have that $K:=\Qs\{ f \, : \, \norma{f} \le 1 \} \subset \{ f \in \B \, : \, \normb{f} \le 1 \}$, thus $K$ is bounded. Using \eqref{eq:hoeldercontinuous} with $n=1$, we deduce that the family $K$ is equicontinuous. Hence, applying the Arzel\`a-Ascoli theorem, $K$ is conditionally compact in $\Cf{\Sp}$ with respect to the topology of uniform convergence, i.e. w.r.t. $\normb{\cdot}$.
\medskip
\Step[3]: Next we show that Assumption \eqref{prop3} holds for $\Qs$, i.e. there exist $k \in \N$ and real numbers $r$ and $R$ with $r < \rho(\Qs)$ and, for all $f \in \B$,
$$ \norma{(\Qs)^k f} \le R \normb{f} + r^k \norma{f}.$$ In particular, $\Qs \in \mathcal{L}(\B,\B)$.
Observe, that it suffices to provide the estimate for \emph{one} $k \in \N$, it is not necessary to prove a geometric decay rate. Since the spectral radius $r(\Qs)=1$, it is enough to show that the inequality holds for some $r' < 1$ in the place of $r^k$, because then $r := {r'}^{1/k}<1$ satisfies the assumption.
Using \eqref{eq:hoeldercontinuous} and the fact that $\Qs$ is a Markov operator, we deduce that for any $n \in \N$,
\begin{align}
\label{eq:normaQs} \norma{(\Qs)^n f} ~&=~ \normb{{(\Qs)^n f}} + \abs{{(\Qs)^n f}}_s \\
~&\le~ \normb{f} + C \normb{f} + D(n) \abs{f}_s ~\le~ (1+C) \normb{f} + D(n) (\abs{f}_s + \normb{f}) \nonumber \\
~&=~ (1+C) \normb{f} + D(n) \norma{f} \nonumber .
\end{align}
But $D(n)$ tends to 0, thus we may choose $k$ such that {\color{black} $D(k) < 1$}, and consequently, Assumption \eqref{prop3} is satisfied with $R:=1 + C$ and $r := D(k)^{1/k} < 1$.
Thus {\color{black} Theorem \ref{thm:ITM}} applies and gives the {\color{black} quasi-compactness of $Q^s$ , i.e. $\B=E \oplus F$ for $Q^s$-invariant closed subspaces $E$ and $F$ with $\dim E <\infty$ and such that $Q^s_{|F}$ has spectral radius strictly smaller than 1, while each eigenvalue of $Q^s_{|E}$ has modulus 1. }
\medskip
\Step[4]: {\color{black} Next we prove that $1$ is a simple eigenvalue}, and the only one of modulus one, i.e. $\dim E=1$.
It is shown in \cite[Theorem 4.13]{BDGM2014}, that for every $f \in \Cf{\Sp}$,
\begin{equation}\label{eq:convQs} \lim_{n \to \infty} (\Qs)^n f = \pis(f). \end{equation}
If now {\color{black}$\Qs f = \lambda f$} with $\abs{\lambda}=1$, then necessarily {\color{black} $\lim_{n \to \infty} \lambda^{n} f \equiv \pis(\lambda f)$}, which implies $\lambda=1$ and $f={\rm const}$.
\medskip
{\color{black} \Step[5]: We infer from Eq. \eqref{eq:normaQs} that
$$ \limsup_{n \to \infty} \rho(Q^s)^{-n} \Big( \sup \, \{\norma{(\Qs)^n f} \, : \, \norma{f} =1 \}\Big) \le 1+C.$$
Therefore, $Q^s$ is {\em quasi compact of diagonal type} in the sense of \cite[Prop. III.1]{HH2001}. Consequently, \cite[Lemma III.3(v)]{HH2001} applies and gives the decomposition
\eqref{eq:decompQs} with $M$ being the projection on $E=\R\1[\Sp]$ with $M(f)=\pis(f)\1[\Sp]$ for all $f \in \B$, and $N:= Q^s -M$.
}
\end{proof}
\subsection{Invertible matrices}
As said before, the ideas of the proofs above were developed by Guivarc'h and Le Page for condition \textit{(i-p)}, the result corresponding to Proposition \ref{prop:Qs} is \cite[Corollary 3.19]{Guivarch2012}. There, the distance $d(x,y)=\abs{x-y}$ for $x,y \in \mathbb{P}^{d-1}$ is the minimal euclidean distance between representants in $\Sd$.
Under assumption \textit{(id)}, a corresponding decomposition, proved in \cite[Proposition 4.3 and Lemma 4.11]{BDMM2013} holds on the (larger) space $\Cf{\Sd}$, for $(X_n)$ is a Doeblin chain under each ${\QQ_x^s}$.
\section{Non-artihmeticity and its consequences}\label{sect:nonarithmetic}
Subsequently, fix $s \in I_\mu$. We will now study implications of the non-arithmeticity and moment assumptions \eqref{eq:non-arithmetic} resp. \eqref{eq:iotamoment} for the family $Q(it)$. {\color{black} Recall from Lemma \ref{lem:FT} the identity}
\begin{equation}
\label{eq:defQit} Q(it)^n f(x) {\color{black} ~=~} \Erw[{\QQ_x^s}]{e^{it S_n} f(X_n)} ~=~ \Erw{q_n^s(x, \mPi_n) e^{it \log \abs{\mPi_n x}} f(\mPi_n \as x)}.
\end{equation}
This section is valid for all types of matrices. Recall that conditions \textit{(i-p)}~and \textit{(id)}~readily imply non-arithmeticity.
Define $\B_\epsilon := \{ f \in \Cf{\S} \, : \, \abs{f}_\epsilon < \infty \}.$
Then we are going to prove the following result:
\begin{thm}\label{thm:prop:nonarithmetic}
Assume that $(X_n, S_n)$ is non-arithmetic under {\color{black} $\Q^s$} or that $\mu$ is non-arithmetic. Assume that \eqref{eq:iotamoment} hold for some $\epsilon >0$. {\color{black} Then $Q(it) \in \mathcal{L}(\B_\epsilon, \B_\epsilon)$ for all $t \in \R$. Moreover, for all $t \neq 0$, the spectral radius $\rho(Q(it)) < 1$ and thus $1-Q(it)$ is invertible in $\mathcal{L}(\B_\epsilon, \B_\epsilon)$.}
\end{thm}
Considering the spectral radius, we have for all $n \in \N$ and all $f \in \Cf{\S}$ that
\begin{equation} \label{eq:Qitf}\abs{Q(it)^n f(x)} \le \E_{{\QQ_x^s}}\abs{e^{itS_n} f(X_n)} = \E_{{\QQ_x^s}} \abs{f(X_n)} = (Q^s)^n \abs{f}(x).\end{equation} Since $\B_\epsilon \subset \Cf{\S}$, this readily shows that $\rho(Q(it)) \le \rho(Q^s)=1$ for all $t \in \R$. To arrive at $\rho(Q(it)) < 1$, the main burden of the proof will be indeed to show that ${\color{black} Q(it)} \in \mathcal{L}(\B_\epsilon,\B_\epsilon)$. This will be done by proving an estimate similar to \eqref{eq:hoeldercontinuous}. The first step in that direction is provided by the following lemma.
\begin{lem}\label{lem:epsilonHoelder}
Let $\ma$ either be an allowable nonnegative matrix or an invertible matrix. Then for all $t \in \R$, $x,y \in \S$ and $0 < \epsilon < 1$, the following estimate holds true:
\begin{equation}
\abs{e^{it \log \abs{\ma x}}- e^{it \log \abs{\ma y}}} \le D \abs{t}^{\epsilon} d\left( x, y \right)^{\epsilon} \left( \frac{\norm{\ma}}{\iota(\ma)} \right)^{\epsilon}
\end{equation}
for some $D > 0$.
\end{lem}
Recall that in the case of nonnegative matrices, the distance $d$ on $\S$ was defined in \ref{subsect:metric}, while $d(x,y)$ equals the minimum of the euclidean distance of representants of $x,y$ from $\Sd$ in the case of nonnegative matrices.
\begin{proof}
We start by noting the some useful inequalities:
\begin{align}\label{ineq1}
\frac12 \abs{e^{it}-e^{is}} = \frac{1}2 \abs{1-e^{i(s-t)}}\le \min\{1, \abs{t-s}\} \le \abs{t-s}^\beta
\end{align}
for all $t,s \in \R$, $\beta \in [0,1]$. Next, for all $a,b > 0$,
\begin{align}\label{ineq2}
\abs{\log a - \log b} = \abs{\int_a^b \, \frac{1}s \, \d s} \le \max\{\frac{1}{a}, \frac1b\} \abs{a-b}.
\end{align}
Finally, if $\ma$ is allowable or invertible, then for all $x \in \S$, $\frac{1}{\abs{\ma x}} \le \frac{1}{\iota(\ma)}$. Putting these inequalities together, we conclude, using Proposition \ref{prop:properties of d} as well in the case of invertible matrices,
\begin{align*}
\abs{e^{it \log \abs{\ma x}}- e^{it \log \abs{\ma y}}} ~&\le~ 2 \abs{t}^{\epsilon} \abs{\log \abs{\ma x} - \log \abs{\ma y}}^{\epsilon} \\
~&\le~ 2 \abs{t}^{\epsilon} \max\{\frac1{\abs{\ma x}}, \frac1{\abs{\ma y}} \}^{\epsilon} \ \abs{\abs{\ma x} - \abs{\ma y}}^{\epsilon} ~\le~ 2 \abs{t}^{\epsilon}\max\{\frac1{\abs{\ma x}}, \frac1{\abs{\ma y}} \}^{\epsilon} \ \norm{\ma}^{\epsilon} \abs{x-y}^{\epsilon} \\
~&\le~ 2 \abs{t}^{\epsilon} C^{-1} d\left( x, y \right)^{\epsilon} \left( \sup_{z \in \S} \frac{\norm{\ma}}{\abs{\ma z}} \right)^{\epsilon} ~\le~ 2 \abs{t}^{\epsilon} C^{-1} d\left( x, y \right)^{\epsilon} \left( \frac{\norm{\ma}}{\iota(\ma)} \right)^{\epsilon}
\end{align*}
\end{proof}
Now we are going to prove an estimate similar to \eqref{eq:hoeldercontinuous}:
\begin{lem}\label{lem:hoelder2}
There is $C >0$ and a sequence $D(n)$ with $\lim_{n \to \infty} D(n) = 0$, such that for all $n \in \N$ and $f \in \B$,
\begin{align}\label{eq:hoeldercontinuous2}
\abs{Q(it)^n f}_\epsilon \le C \normb{f} + D(n) \abs{f}_\epsilon
\end{align}
\end{lem}
\begin{proof}
{\color{black}\begin{eqnarray*}
& & \abs{Q(it)^n f(x) - Q(it)^n f(y)} \\
&= & \abs{ \Erw[{\QQ_x^s}]{e^{it S_n} f(X_n)} - \Erw[{\QQ_y^s}]{e^{itS_n} f(X_n)}} \\
&\le& \E \big( q_n^s(x,\mPi_n) \abs{ f(\mPi_n \as x) - f(\mPi_n \as y) } \big) + \abs{\E \big(q_n^s(x,\mPi_n) - q_n^s(y,\mPi_n) \big) e^{it S_n^x} f(\mPi_n \as y)} \\
& & + \normb{f} \E \big( q_n^s(y,\mPi_n) \abs{e^{it S_n^x} - e^{it S_n^y}} \big) \\ &=& I + II + III.
\end{eqnarray*}}
Similar to the proof of Lemma \ref{lem:hoeldercontinuous}, we obtain the bounds
\begin{align*}
I ~&\le~ \abs{f}_\epsilon d(x,y)^\epsilon \E_{{\QQ_x^s}} c(\mPi_n)^\epsilon =: D(n) \abs{f}_\epsilon d(x,y)^\epsilon, \\
II ~&\le~ \normb{f} C \, d(x,y)^\epsilon
\end{align*}
with $\lim_{n \to \infty} D(n) =0$.
Using Lemma \ref{lem:epsilonHoelder}, we deduce
\begin{align*}
III ~&\le~ D \normb{f} \abs{t}^{\epsilon} d\left( x, y \right)^{\epsilon} \E_{{\QQ_y^s}}\left( \frac{\norm{\mPi_n}}{\iota(\mPi_n)} \right)^{\epsilon} \\
~&\le~ D' \normb{f} \abs{t}^{\epsilon} d\left( x, y \right)^{\epsilon} \Erw{ \norm{\mPi_n}^{s+\epsilon} {\iota(\mPi_n)}^{-\epsilon}} \\
~&\le~ D' \normb{f} \abs{t}^{\epsilon} d\left( x, y \right)^{\epsilon} \left(\E{ \norm{\mA_1}^{s+\epsilon} {\iota(\mA_1)}^{-\epsilon}}\right)^n,
\end{align*}
where the last expression is finite due to assumption \eqref{eq:iotamoment}.
\end{proof}
\medskip
\begin{proof}[Proof of Theorem \ref{thm:prop:nonarithmetic}]
Lemma \ref{lem:hoelder2} together with \eqref{eq:Qitf} proves that $Q(it)$ is a self-map of $\B_\epsilon$. Since {\color{black} $\rho(Q(it)) \le 1$}, it remains to exclude the possibility $\rho(Q(it))=1$, which we will do by contradiction.
Assuming that the spectral radius $\rho(Q(it))=1$, one can proceed as in Section \ref{sect:quasi-compact} in order to show that the Ionescu-Tulcea-Marinescu theorem applies for $Q=Q(it)$ with
$$ \normb{f}:= \sup_{x \in \S} \abs{f(x)}, \qquad \norma{f}_\epsilon:=\normb{f}+\abs{f}_\epsilon$$
and the Banach space
$$ \B_\epsilon := \{ f \in \Cf{\S} \, : \, \abs{f}_\epsilon < \infty \} = \{ f \in \Cf{\S} \, : \, \norma{f}_\epsilon < \infty \}$$
equipped with the norm $\norma{\cdot}_\epsilon$.
The theorem yields that there has to be an eigenvalue with modulus equal to the spectral radius of $Q(it)$, i.e. with modulus equal to 1.
Hence, suppose there is an eigenfunction $f$ such that $Q(it) f = \lambda f$ with $\abs{\lambda}=1$. {\color{black} Let $x_0 \in \Sp$ be such that $\abs{f(x_0)}=[f]$. Then \eqref{eq:Qitf} implies that $\abs{f}(x_0) \le ((Q^s)^n \abs{f})(x_0)$ and hence by \eqref{eq:convQs}, $\abs{f}(x_0) \le \pis(\abs{f})$. But the right hand side is a convex combination of $(\abs{f}(x))_{x \in V(\Gamma)}$ (see Proposition \ref{prop:transferoperators}, \eqref{suppnus}). Consequently, $\abs{f}$ has to be constant on $V(\Gamma)$.}
Thus, we can assume that $f(x)=e^{i \vartheta(x)}$ on $V(\Gamma)$ for a continuous function $\vartheta: \S \to \R$. Consequently,
$$ \E_{{\QQ_x^s}} e^{i t S_1 + i \vartheta(X_1)} = e^{i \theta + i \vartheta(x)}.$$
But this contradicts the non-arithmeticity of $(X_n, S_n)$ under {\color{black} $\Q^s$}; and, since $\vartheta$ is continuous, as well the nonarithmeticity of $\mu$, using Lemma \ref{lem:implications_arithmetic}.
\end{proof}
\noindent {\bf Remark}. Observe, that we only did prove the estimate
$$ \normb{Q(it)^n f}_\epsilon \le R \norma{f} + D(n) \normb{f}_\epsilon, $$
with $D(n)$ tending to 0. From this, property \eqref{prop3} of the Ionescu-Tulcea-Marinescu theorem can be deduced only if $\rho(Q(it))=1$, since otherwise, we do not know whether $r := D(n)^{1/n} < \rho(Q(it))$ holds for some $n$, since we do not know the rate of convergence for $D(n) \to 0$. So we do not know yet whether $Q(it)$ is quasi-compact for $t \neq 0$. Nevertheless, for small $t$, quasi-compactness will follow from the perturbation theorem below.
\section{The Perturbation Theorem}\label{sect:perturbation}
This section as well is valid for nonnegative and invertible matrices.
Recall {\color{black} from Lemma \ref{lem:FT} the fundamental identity
$$ \phi_{n,x}(t) ~=~ \E_{\Q_x^s}\left( e^{itS_n}\right) ~=~ Q(it)^n \1[\S](x).$$}
In this section, we are going to apply an holomorphic perturbation theorem for $\Qs$ in order to show that (for small t)
$$ Q(it)^n = \theta(it)^n M(it) + N^n(it),$$
where $M(it)$ is a rank one-projection, which commutes with $N^n(it)$, and $\rho(N(it)) < \rho(Q(it))$. Using this decomposition, we will be -- roughly speaking -- able to replace
$$\phi_{n,x}(t/\sqrt{n}) \approx \theta(t\sqrt{n})$$ for large $n$ in the proof of the Edgeworth expansion.
Fix the parameter $s \in \interior{I_\mu}$ as well as $\epsilon$ such that \eqref{eq:iotamoment} is satisfied. By what has been shown above, $\Qs \in \mathcal{L}(\B_\epsilon, \B_\epsilon)$ is quasi-compact with a simple dominant eigenvalue $1$. This, and the holomorphicity of the mapping $z \mapsto Q(z)$, shown below, will be the main ingredients for the application of a perturbation theorem.
\subsection{Perturbation theory for $\Qs$}
\begin{lem}\label{lem:defQz}
Choose $\delta>0$ such that $(s-\delta, s+ \delta) \subset I_\mu$ and $s+\delta > \epsilon$. Then for all
$z\in H_\delta:=\{z \in \C \, : \, \Re z \in (-\delta, \delta) \}$, the operator $Q(z)$ on $\B_\epsilon$, which is given by
$$Q(z) f(x) := \frac{1}{\es(x) k(s)} \int \abs{\ma x}^{s +z} f(\ma \as x) \es(\ma \as x) \, \mu(\d \ma) = \E_{\QQ_x^s} \big[e^{z S_1} f(X_1)\big],$$
is well defined. The mapping $Q : H_\delta \to \mathcal{L}(\B_\varepsilon,\B_\varepsilon)$, $z \mapsto Q(z)$ is holomorphic.
\end{lem}
\begin{proof}
Recalling that $\es$ is bounded from below and above, it follows that
\begin{equation}\label{eq:boundperturbation} [Q(z) f] \le K \E \norm{\mA}^{s + \Re z} < \infty, \end{equation}
since $s + \Re z \in I_\mu$.
Together with Lemma \ref{lem:hoelder2}, this proves that $Q(z) \in \mathcal{L}(\B_\epsilon,\B_\epsilon)$.
Now we can show that $z \mapsto Q(z)$ is weakly holomorphic, i.e. for any $f \in \B_\varepsilon$, $\nu \in \B_\varepsilon'$ ({\color{black} the dual space of $\B_\varepsilon$}), {\color{black} $z \mapsto \int Q(z) f \d \nu$} is holomorphic. This readily implies that $z \mapsto Q(z)$ is (strongly) holomorphic, see \cite[Theorem V.3.1]{Yosida1980}.
In order to show weak holomorphicity, consider arbitrary $f, \nu$ and a closed curve $\gamma \subset B_\delta(0) \subset \C$. Then
\begin{align*}
&~\int_\gamma \left( \int_{\S} (Q(z) f)(x) \, \nu (\d x) \right) \ \d z \\
=&~ \int_{\S} \left( \frac{1}{\es(x) k(s)} \int f(\ma \as x) \es(\ma \as x) \ \left\{\int_{\gamma} e^{(s+z) \ln \abs{\ma x}} \ \d z \right\} \ \mu(\d \ma) \right) \ \nu(\d x)
=0,
\end{align*}
for the innermost function is holomorphic in $z$. The change of the order of integration is guaranteed by the estimate \eqref{eq:boundperturbation}.
\end{proof}
Now we can apply the following perturbation theorem \cite[Theorem III.8]{HH2001}.
\begin{thm}\label{theorem:perturbation}
Let $G_0:=B_\delta(0) \subset \C$ and let $(Q(z))_{z \in G_0}$ be a collection of elements of $\mathcal{L}(\B_\varepsilon,\B_\varepsilon)$ such that
\begin{itemize}
\item[(H1)] $z \mapsto Q(z)$ is holomorphic on $G_0$,
\item[(H2)] $Q(0)$ has one dominating simple eigenvalue and $\rho(Q(0))=1$.
\end{itemize}
Then there exist $G_1:=B_{\delta_1}(0) \subset \C$, $G_1 \subset G_0$ and holomorphic mappings
$$ \theta : \, G_1 \to \C, \quad r :\, G_1 \to \B_\varepsilon, \quad \nu :\, G_1 \to \B_\varepsilon', \quad N: \, G_1 \to \mathcal{L}(\B_\varepsilon,\B_\varepsilon) $$
such that for all $n \ge 1$, $z \in G_1$
$$ Q^n(z)
= \theta(z)^n M(z) + L(z)^n, $$
with $Q(z) r(z) = \theta(z) r(z)$ and $\nu(z) Q(z) = \theta(z) \nu(z)$.
Moreover, {\color{black} for each $l_0 \in \N$} there exist constants $\eta_1, \eta_2 >0$, $c \ge 0$ such that for all $z \in G_1$,
$$ \abs{\theta(z)} \ge 1 - \eta_1, \text{ and } \max\left\{ \norm{\frac{\d^l}{\d z^l} L(z)^n} \, : \, {\color{black} l \le l_0}\right\} \le c (1 - \eta_1 - \eta_2)^n.$$
\end{thm}
\subsection{The operators $R(t)$}
For the Edgeworth expansion, we will consider a slightly different operator, namely such that $S_1$ becomes centered: Let $q:= \E_{\Q_s} S_1$ denote the stationary drift of $S_1$ under $\Q_s$, and define the family $(R(t))_{t \in\R}$ of operators by
\begin{equation}\label{def:operator R}
R(t) f(x) := e^{-itq} Q(it)f(x) = \Erw[{{\QQ_x^s}}]{e^{it(S_1 -q)} f(X_1)}.
\end{equation}
Upon defining
\begin{equation}
\lambda(t) := e^{-itq} \theta(it), \qquad N(t):= e^{-itq} L(it), \qquad \Pi(t)
= M(it),
\end{equation}
we obtain the following corollary {\color{black} of Lemma \ref{lem:defQz} and Theorem \ref{theorem:perturbation}}.
\begin{cor}\label{cor:Rt}
{\color{black} There is $\delta_1 >0$} such that for all $t \in G:= (- \delta_1, \delta_1)$,
$$ R(t)^n = \lambda(t)^n \Pi(t) + N(t)^n,$$
with $\Pi(t) N(t) = N(t) \Pi(t) =0$. {\color{black} For each $l_0 \in \N$ there is $\eta=\eta(l_0) >0$ and $c=c(l_0) < \infty$} such that \begin{equation} \label{eq:boundN} \max\left\{ \norm{\frac{\d^l}{\d z^l} N(t)^n} \, : \, {\color{black} l \le l_0} \right\} \le c (1 - \eta)^n. \end{equation}
The mappings $\lambda: G \to \C$, $\Pi : G \to \mathcal{L}(\B_\varepsilon,\B_\varepsilon)$ and $N: G \to \mathcal{L}(\B_\varepsilon,\B_\varepsilon)$ are $\mathcal{C}^\infty$, the latter ones in the strong operator sense.
\end{cor}
{\color{black} For all purposes below, we can choose $l_0=3$, and may therefore consider $\eta=\eta(3)$, $c=c(3)$ fixed.}
In order to prove the Edgeworth expansion, we will make as well use of the following result{\color{black}, which is inspired by \cite[Lemma 3.19]{BDG2010}.}
\begin{lem}\label{missing estimate}
Let $K \subset \R \setminus \{0\}$ be compact. Then for each $f \in \B_\epsilon$, there is $\rho < 1$ such that for all $t \in K$
\begin{equation} [R(t)^n f] \le \rho^n [f]. \end{equation}
\end{lem}
\begin{proof}
Fix $f \in \B_\varepsilon$. For each $n \in \N$, the mapping
$$t \mapsto [R(t)^n f]^{1/n} = \left( \sup_{x \in S} \abs{\Erw{q_n^s(x,\mPi_n) e^{it(S_n^x -q)} f(X_n^x)}} \right)^{1/n}$$
is continuous. Hence, $\rho_f(t):= \limsup_{n \to \infty} [R(t)^n f]^{1/n} $ is upper semicontinuous, thus it attains it maximum on the compact set $K$, in $t_0 \neq 0$, say.
But $\rho_f(t_0) \le \rho(R(t_0)) = \rho(Q(t_0)) < 1$, hence the assertion follows.
\end{proof}
\section{Taylor expansion of $\lambda(t)$ and positivity of the asymptotic variance}\label{sect:taylor}
In this section, which is valid for all types of matrices, we are going to relate the first and second order coefficients of the Taylor expansion of $\lambda$ with the expectation (which equals zero in fact) resp. the asymptotic variance of $S_n-nq$ under ${\QQ_x^s}$. Moreover, we are able to prove that the asymptotic variance is positive as soon as $\mu$ is non-arithmetic.
\begin{lem}
\label{lem: lambda}\label{lem:sigma1}
Assume that $\mu$ satisfies $\condC$, \textit{(i-p)}~or \textit{(id)}, and that $s \in \interior{I_\mu}$.
Then there is $\sigma \ge 0$ and $m_3 \in \R$ such that
$${\color{black}
\lambda(t) = 1 - \frac{\sigma^2}2\, t^2 -i \frac{m_3}{6}\, t^3 + o(t^3),}
$$ and
\begin{equation}
\label{eq: sigma}
\sigma^2 = \lim_{n \to \infty} \frac{1}{n} \E_{\QQ^s} (S_n-nq)^2 \qquad
m_3= \lim_{n \to \infty} \frac{1}{n} \E_{\QQ^s} (S_n-nq)^3.
\end{equation}
For each $x\in \S$, the value
\begin{equation}
\label{eq: bx}
b(x) = \lim_{n \to \infty} \E_{\QQ_x^s} (S_n-nq)
\end{equation}
is well defined, and the mapping $b \in \B_\varepsilon$.
It holds that
\begin{equation}\label{eq:bx}
b(x) = \Erw[{\QQ_x^s}]{(S_1-q) {\color{black} +} b(X_1) }
\end{equation}
Moreover, \begin{equation}
{\color{black} \sigma^2 = \E_{\QQ^s} \left[ \Bigl( (S_1 - q) + b(X_1) \Bigr)^2 - b(X_1)^2 \right],} \end{equation}
and
\begin{equation}
\sup_{n \in \N} \abs{n\sigma^2 - \E_{\QQ^s}(S_n - nq)^2} < \infty. \label{eq:sn2finite}
\end{equation}
\end{lem}
To prove Lemma \ref{lem: lambda} we reason as in \cite[Lemma 8.3 \& Lemma 8.4 ]{Herve2010}.
\begin{proof}[Proof of Lemma \ref{lem: lambda}]
{\sc Step 1.} First we prove that $\lambda'(0)=0$.
Differentiating the equation
$
R(t)\Pi(t)\1 = \lambda(t)\Pi(t)\1
$ in the operator sense and computing its value at 0, we obtain
\begin{equation}
\label{eq: 1}
{\color{black} R'(0)\1 + R(0)\Pi'(0)\1 = \lambda'(0)\1 + \Pi'(0)\1.}
\end{equation}
Both sides of the above equation are bounded continuous functions, so computing their integral with respect to the measure $\pi$ we obtain
$$\pi(R'(0)\1) +\pi(\Pi'(0)\1) = \lambda'(0) + \pi(\Pi'(0)\1)
$$
we have
$$
{\color{black} \lambda'(0) = \pi(R'(0)\1) = i\E_{\Q^s}[S_1-q]=0.}
$$
{\sc Step 2. } Now we justify, that the function $b(x)$ is well defined as a function in {\color{black} $\B_\epsilon$}.
{\color{black} Observe first that by Lemma \ref{lem:FT} with $z=0$ and \eqref{eq:qns}, we have
\begin{align}
(Q^s)^{n} f(x) ~&=~ \E_{{\QQ_x^s}}[f(X_n)] \nonumber \\
&=~ \E\, q_n^s(x, \mPi_n) \int_{\Gamma} q_1^s(\mPi_n \as x, \ma) \big( \log \abs{\ma (\mPi_n \as x)} -q \big) \, \mu(d\ma) \nonumber \\
&=~ \E \, \int_{\Gamma} q_{n+1}^s(x, \ma \mPi_n) \big( \log \abs{\ma \mPi_n x)} - \log \abs{ \mPi_n x} -q \big) \, \mu(d\ma) \nonumber \\
&=~ \E_{\Q_x^s} \left( S_{n+1} - S_n -q \right)~=~ \E_{{\QQ_x^s}}[S_{n+1}] - \E_{{\QQ_x^s}}[S_n] -q. \label{eq: 2}
\end{align}
}
Next by \eqref{eq: 1} for any $k$ we have (recall $R(0)=Q^s$)
\begin{equation}\label{eq:3}
i(Q^s)^k \E_{{\QQ_x^s}} [S_1-q] + (Q^s)^{k+1} \Pi'(0)\1(x) = (Q^s)^k \Pi'(0)\1(x).
\end{equation}
Hence summing over $k=0,1,\ldots, n-1$ we obtain
$$
i\sum_{k=0}^{ n-1} (Q^s)^{k}\E_{{\QQ_x^s}}[S_1-q] +
\sum_{k=0}^{n-1} (Q^s)^{k+1} \Pi'(0)\1(x) = \sum_{k=0}^{n-1} (Q^s)^k \Pi'(0)\1(x).
$$
Thus by \eqref{eq: 2}
$$i\E_{{\QQ_x^s}} [S_n-nq] +(Q^s)^n \Pi'(0)\1(x) = \Pi'(0)\1(x).
$$
Since {\color{black} $\Pi'(0)\1\in \B_\epsilon \subset \B$}, the limit $(Q^s)^n \Pi'(0)\1(x)$ exists by Proposition \ref{prop:Qs} and is equal to $\pi(\Pi'(0)\1)$. Deriving the equation $\Pi^2(t)=\Pi(t)$ and computing the result at 0 we obtain $$\pi(\Pi'(0)\1)=0.$$ Thus, the limit
$\lim\E_{{\QQ_x^s}}[S_n-nq]$ exists, equals
\begin{equation}\label{eq:defbx} {\color{black} b(x)~:=~ \frac{1}{i}\, \Pi'(0)\1(x)},
\end{equation} and thus $b$ is well defined and is an element of $\B_\varepsilon$, since $\Pi'(0)$ maps $\1$ into $\B_\varepsilon$.
The formula \eqref{eq:bx} follows from \eqref{eq:3} for $k=0$.
{\sc Step 3.} Using the above, we obtain the following Taylor expansions, valid for small $t$:
$$ \lambda(t)^n = 1 + n \lambda''(0) \frac{t^2}{2} + n \lambda^{(3)} \frac{t^3}{6} + o(t^3),$$
$$ \pis\left( \Pi(t) \1 \right) = 1 + d_1 \frac{t^2}{2} + d_2 \frac{t^3}{6} + o(t^3);$$
as well as the classical expansion for the characteristic function, i.e.
$$ \E_{\QQ^s} e^{it(S_n - nq)} ~=~ 1 - \E_{\QQ^s}(S_n - nq)^2 \, \frac{t^2}{2} - i \, \E_{\QQ^s} (S_n - nq)^3 \, \frac{t^3}6 + o(t^3).$$
From the fundamental identity,
$$ \E_{\QQ^s} e^{it(S_n - nq)} ~=~ \pis(R(t)^n \1) = \lambda(t)^n \pis(\Pi(t)\1) + \pis(N(t)^n \1), $$
using {\color{black} the bounds \eqref{eq:boundN} (with $l_0=3$) for $N$ as well, }
we deduce that
$$ n \lambda''(0) + d_1 + O((1-\eta)^n) ={\color{black} - \E_{\QQ^s}(S_n - nq)^2} $$
and
$$ n \lambda^{(3)}(0) + d_2 + O((1-\eta)^n) = {\color{black} -i\,\E_{\QQ^s}(S_n - nq)^3} .$$
Hence, the identification of $\sigma^2$ and $m_3$ as well as the boundedness assertion follow.
\Step[4]: Finally, we provide the formula for $\sigma^2$. Differentiating $R(t)\Pi(t)\1 = \lambda(t)\Pi(t)\1$ twice and integrating against $\pis$, using {\color{black} $\pis R(0) = \pis$}, we obtain
$$ \pis(R''(0)\1) + 2\, \pis(R'(0) \Pi'(0)\1) = \lambda''(0),$$
hence recalling from above that {\color{black} $i \,b(x) = \Pi'(0)\1(x)$},
$$ {\color{black} - }\int \E_{{\QQ_x^s}} (\log \norm{\mA_1 x}-q)^2 \, \pis(\d x) \, {\color{black} - } \, 2\, \int \E_{{\QQ_x^s}} \left[ \log (\norm{\mA_1 x}-q) \, b(\mA_1 \as x) \right] \, \pis(\d x) ~=~ \lambda''(0),$$
i.e.
$$ \sigma^2 = \E_{\QQ^s} \left[ (S_1 - q)^2 + 2 (S_1-q) b(X_1) \right]$$
and the result follows by quadratic extension inside the expectation.
\end{proof}
Using the above formula for $\sigma^2$, one can show that non-arithmeticity readily implies that $\sigma^2 >0$.
\begin{lem}\label{cor:arithmetic}
Assume that $\sigma^2 =0$, then
$$ S_1 ~=~ q - b(X_1) + b(X_0) \qquad \QQ^s\text{-a.s.}, $$
in particular, $(X_n, S_n)$ is arithmetic under $\QQ^s$ and $\mu$ is arithmetic.
\end{lem}
\begin{proof}
If $\sigma^2=0$, then it follows from \eqref{eq:sn2finite}, that
\begin{align*} \int b(x)^2 \, \pis(\d x) =&~ \int \left[\lim_{n \to \infty} \E_{{\QQ_x^s}}(S_n -nq) \right]^2 \pis(\d x)\le \int \liminf_{n \to \infty} \E_{{\QQ_x^s}}(S_n -nq)^2 \pis(\d x) \\
\le&~ \liminf_{n \to \infty} \int \E_{{\QQ_x^s}}(S_n -nq)^2 \pis(\d x) \le \sup_{n \in \N} \E_{\QQ^s} (S_n - nq)^2 < \infty.
\end{align*}
Then we may rewrite the formula from Lemma \ref{lem:sigma1} to read
\begin{align}\label{eq:5}
\sigma^2 ~=&~ \E_{\QQ^s} \Bigl( (S_1 - q) + b(X_1) \Bigr)^2 - \E_{\QQ^s} b(X_1)^2 = \E_{\QQ^s} \Bigl( (S_1 - q) + b(X_1) \Bigr)^2 - \E_{\QQ^s} b(X_0)^2.
\end{align}
Using \eqref{eq:bx}, we see that
$$ \Erw[{\QQ^s}]{\Bigl((S_1-q)+b(X_1)\Bigr)b(X_0)} = \int \, b(x) \Erw[{{\QQ_x^s}}]{\Bigl((S_1-q)+b(X_1)\Bigr)} \, \pis(\d x) = \int \, b(x)^2 \pis(\d x),
$$
which we use in \eqref{eq:5} to obtain (through binomial formula) that
$$ 0 = \sigma^2 = \E_{\QQ^s} \Bigl( (S_1 - q) + b(X_1) - b(X_0) \Bigr)^2 = \int_{V(\Gamma)} \pis(\d x) \, \int_{\supp \, \mu} \mu(\d \ma) \, \Bigl( \log \abs{\ma x} - q + b(\ma \as x) - b(x) \Bigr)^2. $$
This gives the assertion; and the arithmeticity of $\mu$ follows, since the function $b$ is continuous (see Lemma \ref{lem:implications_arithmetic}).
\end{proof}
Finally, we note some expressions for derivatives of $k$.
\begin{cor}\label{cor:k}
The function $k(s)$ is $\mathcal{C}^\infty$ on $\interior{I_\mu}$, and
$$ \frac{k'(s)}{k(s)}=q=\E_{\Q^s} S_1, \qquad \frac{k^{(2)}(s)}{k(s)} = q^2 +\sigma^2. $$
\end{cor}
\begin{proof}
Recalling the definition of $\Ps$, we see that for $\epsilon \in (- \delta_1, \delta_1)$,
\begin{align*} (\Ps[s+\epsilon])^n f(x) &~= \es(x) k(s)^n Q(\epsilon) \frac{f}{\es} (x) \\
&~= k(s)^n \theta(\epsilon)^n \es(x) r(\epsilon)(x) \int_{\Sp} f(y)/ \es(y) \, \nu(\epsilon)(dy) + \es(x) k(s)^n N(\epsilon) \frac{f}{\es}(x)
\end{align*}
By Proposition \ref{prop:transferoperators}, $\Ps[s+\epsilon]$ has a unique strictly positive eigenfunction, which is then given by $r_s(x)r(\epsilon)(x)$ and thus the corresponding eigenvalue
equals
\begin{equation}\label{eq:kappatheta} k(s + \epsilon)= k(s) \theta(\epsilon).\end{equation}
{\color{black} By Theorem \ref{theorem:perturbation}, the function $\theta$ is holomorphic in a neighbourhood of $0$, hence $\mathcal{C}^\infty$ in $0$ and so is $k$, with $ k^{(n)}(s) = k(s) \theta^{(n)}(0)$}. Recalling that $\lambda(t)=e^{-itq}\theta(it)$, we obtain
$$ \lambda'(0)=i \frac{k'(s)}{k(s)} - iq, \qquad \lambda^{(2)}(0) = -q^2 + 2 q \theta'(0) - \theta^{(2)}(0) = - q^2 + 2q \frac{k'(s)}{k(s)} - \frac{k^{(2)}(s)}{k(s)}.$$
Since $\lambda'(0)=0$, the assertions follow.
\end{proof}
\section{The Edgeworth expansion}\label{sect:edgeworth}
In this section we are going to prove a third-order Edgeworth expansion for $S_n$ w.r.t. the measure ${\QQ_x^s}$, valid for all types of matrices.
We fix real $s \in \interior{I_\mu}$, and denote by $q:= \E_{\QQ^s} S_1$ the stationary drift of $(S_n)_{n \in \N}$. We will use the operator
$R(t) f(x) = \E_{\QQ_x^s} [ e^{it (S_1 - q)} f(X_1)].$
Let
$$ F_{n,x}(t) := {\QQ_x^s} \left\{ \frac{S_n-nq}{\sqrt{n \sigma^2}} \le t \right\}.$$
be the cumulative distribution function {\color{black} of the standardized version of $S_n$, and write $\Phi$ for the cumulative distribution function of the standard normal distribution}. Then we have the following result.
\begin{thm}\label{thm:edgeworth}
Assume that $\mu$ satisfies $\condC$ and is non-arithmetic, or that \textit{(i-p)}~ or \textit{(id)}~ holds. Assume moreover that \eqref{eq:iotamoment} holds for some $\epsilon >0$. Then
$$ \lim_{n \to \infty} \, \sup_{x \in \Sp} \, \left\{ \sqrt{n} \ \sup_{u \in \R} \abs{ F_{n,x}(u) - \Phi(u) - \frac{m_3}{6 \sigma^3 \sqrt{n}}(1-u^2) \phi(u) + \frac{b(x)}{\sigma \sqrt{n}} \phi(u) } \right\} =0,$$
for quantities $b(x) \in \R$, $\sigma^2 > 0$, $m_3 \in \R$ as defined in \eqref{eq: sigma} and \eqref{eq: bx}.
\end{thm}
\begin{proof}[Proof of Theorem \ref{thm:edgeworth}] We proceed as in \cite{Herve2010}, i.e. we try to follow the standard proof
as in the i.i.d. case. Recall that, since we assume non-arithmeticity, $\sigma^2>0$ due to Corollary \ref{cor:arithmetic}.
{\sc Step 1}. We define the function
$$
{\color{black} G_n(u) ~:=~ \Phi(u) + \frac{m_3}{6\sigma^3 \sqrt n} (1-u^2)\phi(u) - \frac{b(x)}{\sigma \sqrt n}\phi(u) ~=~ \Phi(u) - \frac{m_3}{6\sigma^3 \sqrt n} \phi''(u) - \frac{b(x)}{\sigma \sqrt n}\phi(u),\qquad u\in \R.}
$$ Here $\Phi$ denotes the {\color{black} cumulative distribution function}, and $\phi$ the densitiy function of a standard normal distribution.
One can easily see that the derivative of $G_n$,
$$ {\color{black} G_n'(u) ~=~ \phi(u) - \frac{m_3}{6 \sigma^3 \sqrt{n}} \phi^{(3)}(u) - \frac{b(x)}{\sqrt{n}} \phi'(u) }$$
has exponential decay both at $+\infty$ and $-\infty$, uniformly in $n$.
Let {\color{black} $\gamma_n(t) := \int e^{itu}\, G_n'(u)\, du$ } be the Fourier transform of $G_n'$, then
$$
\gamma_n(t) = \bigg( 1+\frac{m_3}{6 \sigma^3 \sqrt n}(it)^3 \bigg) \cdot e^{-\frac 12 t^2} + \bigg( it \frac{b(x)}{\sigma \sqrt n} \bigg) e^{-\frac 12 t^2}
$$
Denote
\begin{eqnarray*}
\gamma_{0,n}(t) &:=& \bigg( 1+\frac{m_3}{6\sigma^3 \sqrt n}(it)^3 \bigg) \cdot e^{-\frac 12 t^2}, \\
\gamma_{x,n}(t) &:=& \bigg( it \frac{b(x)}{\sigma\sqrt n} \bigg) e^{-\frac 12 t^2}, \\
\varphi_{n,x}(t) &:=& (R(t))^n(\1)(x) = \E_{{\QQ_x^s}}[ e^{it (S_n-nq)}], \\
{\color{black} m } &:=&{\color{black} \sup_{n \in \N} \sup_{u \in \R} \abs{G_n'(u)} < \infty.}
\end{eqnarray*}
By the Berry-Essen inequality (see \cite[XVI.(3.13)]{Feller1971})
we have that for all $T >0$,
\begin{equation}
\label{eq: BE}
\sup_{u\in \R} \big| F_{n,x}(u) - G_n(u) \big| \le \frac 1{\pi} \int_{-T}^T \bigg|\frac{\varphi_{n,x}(\frac t{\sigma\sqrt n}) -\gamma_n(t) }{t}\bigg|dt + \frac{24 m}{\pi T}.
\end{equation}
Next, fix $\varepsilon >0 $, choose $a$ such that $\frac{24 m}{\pi a}<\varepsilon $. Then with $T=a \sqrt{n}$, $\frac{24 m }{\pi T}\le \frac{\varepsilon}{\sqrt n}$.
We choose $\delta< \min\{a,\delta_1\}$, where $\delta_1$ is given by Corollary \ref{cor:Rt}, i.e. for $t \in (-\delta, \delta)$, the perturbation theory for $R(t)$ holds.
Now we want to estimate the integral in $\eqref{eq: BE}$ by $O(\frac{\varepsilon}{\sqrt n})$. For this purpose we divide the integral into two parts
\begin{eqnarray*}
A_n &=& \int_{\sigma \delta \sqrt n\le |t| \le \sigma a \sqrt n}\bigg|\frac{\varphi_{n,x}(\frac t{\sigma \sqrt n}) -\gamma_n(t) }{t}\bigg|dt,\\
B_n &=& \int_{|t|\le \sigma\delta \sqrt n} \bigg|\frac{\varphi_{n,x}(\frac t{\sigma \sqrt n}) -\gamma_n(t) }{t}\bigg|dt.
\end{eqnarray*}
{\sc Step 2.} We prove that $A_n\le \frac{\varepsilon}{\sqrt n}$ for appropriately large $n$. By Lemma \ref{missing estimate}, we have for $u$ such that $\delta<|u|<a$ and all $x$ the estimate $|\varphi_{n,x}(u)| = |(R(u))^n(\1)(x) | \le \rho^n$, hence
$$
\int_{\sigma\delta \sqrt n \le |t| \le \sigma a\sqrt n}\frac{\varphi_{n,x}(\frac t{\sigma \sqrt n})}{|t|}dt =
\int_{\delta \le |u| \le a}\frac{\varphi_{n,x}(u)}{|u|}du \le C(a,\delta) \rho ^n.
$$ Moreover
$$
\int_{\sigma \delta \sqrt n \le |t| \le \sigma a\sqrt n} \frac{|\gamma_n(t)|}{|t|}dt \le C e^{-\sqrt n}.
$$
{\sc Step 3. } Now we estimate the last term $B_n$ to be smaller than $\frac{\varepsilon}{\sqrt n}$. By Corollary \ref{cor:Rt} we write for $\frac{|t|}{\sigma \sqrt n}<\delta$ (recall that for such small values, the perturbation theory applies) \begin{eqnarray*}
\varphi_{n,x}\Big( \frac t{\sigma \sqrt n}\Big) - \gamma_n(t) &=& \lambda^n\Big( \frac t{\sigma \sqrt n} \Big) \Pi\Big( \frac t{\sigma \sqrt n} \Big) \1(x)
+N^n \Big( \frac t{\sigma \sqrt n} \Big) \1(x) - \gamma_{0,n}(t) - \gamma_{x,n}(t)\\
&=& \bigg( \lambda^n\Big( \frac t{\sigma \sqrt n} \Big) -\gamma_{0,n}(t) \bigg) + \lambda^n\Big( \frac t{\sigma\sqrt n} \Big)\bigg(
\Pi\Big( \frac t{\sigma \sqrt n} \Big) \1(x) - 1 - i t \frac{b(x)}{\sigma \sqrt n}
\bigg)\\ &&+ it \frac{b(x)}{\sigma \sqrt n}\bigg( \lambda^n\Big( \frac t{\sigma \sqrt n} \Big) -e^{-\frac 12 t^2} \bigg)
+ N^n\Big( \frac t{\sigma \sqrt n}\Big)\1(x)\\
&=& I_1(t) + I_2(t,x) + I_3(t,x)+ I_4(t,x).
\end{eqnarray*}
Thus, we have to estimate four expressions. {\color{black} For this purpose we will use the Taylor expansion
$$\lambda(u) ~=~ 1 - \frac{\sigma^2}{2} u^2 - i \frac{m^3}{6}u^3 + o(u^3),$$ given in Lemma \ref{lem: lambda}. The function $f(u)=\log \lambda(u) + \frac{\sigma^2}{2}u^2$ then satisfies $f(0)= f'(0)=f''(0)=0$ and $f^{(3)}(0)=-im_3$ and hence
$$ n f\big(\frac{t}{\sigma\sqrt{n}} \big) ~=~ -i \frac{m_3 t^3}{6 \sigma^3 \sqrt{n}} + o(t^3/\sqrt{n}). $$
Moreover, by choosing $\delta$ small enough (but fixed!), we can achieve that for all $u \in (-\delta, \delta)$,
$$ \abs{f(u)} \le \frac{1}{4} u^2 \quad \text{ and } \quad \abs{\frac{m_3}{6}u^3} \le \frac{1}{4} u^2, \quad \text{hence} \quad \max\left\{n \big|f (\frac{t}{\sigma\sqrt{n}}) \big|, \frac{m_3 t^3}{6 \sigma^3 \sqrt{n}} \right\} ~\le~\frac{1}{4}t^2.$$
In particular, with this choice of $\delta$, $\abs{\lambda^n(t/(\sigma\sqrt{n}))} \le e^{-\frac{1}{4}t^2}.$
Considering now $I_1(t)$, we obtain, using the inequality
\begin{equation}\label{eq:feller}\abs{e^u -1 - v} \le (\abs{u-v} + \frac{1}{2} \abs{v}^2 ) e^{\max(\abs{u}, \abs{v})},
\end{equation} which is valid for all $u, v \in \C$ (see \cite[XVI.(2.8)]{Feller1971}),
$$ \abs{I_1(t)} = e^{-\frac{1}{2} t^2} \abs{\exp \left( n f\big(\frac{t}{\sigma\sqrt{n}} \big) \right) - 1 + i \frac{m_3 t^3}{6\sigma^3 \sqrt{n} }} ~\le~ e^{-\frac{1}{2}t^2} \left( t^3 o(\frac{1}{\sqrt{n}}) + t^6 O(\frac{1}{n}) \right) e^{\frac{1}{4} t^2}, $$
from which we infer that
$$ \int_{\abs{t} \le \sigma \delta \sqrt{n}} \, \abs{\frac{I(t)}{t}} dt ~\le~ \left(2 \int_0^\infty (t^3 + t^6) \, e^{-\frac{1}{4} t^2} dt \right) \cdot o\big( 1/ \sqrt{n} \big) $$
To estimate the integral of $I_2(t)$ we use
the bound on $\lambda$ from above, a second order Taylor expansion for $\Pi(t)$ (cf. Corollary \ref{cor:Rt}) and that $\Pi'(0)\1(x) = ib(x)$ (see \eqref{eq:defbx}). Then
$$ \abs{I_2(t,x)} \le e^{-\frac{1}{4} t^2} \abs{\Pi(0)\1(x) + \frac{t}{\sigma \sqrt{n}}\Pi'(0)\1(x) + O(\frac{t^2}{ n}) - 1 - it \frac{b(x)}{\sigma \sqrt{n}}} $$
and consequently
$$ \int_{|t|\le \sigma \delta \sqrt n} I_2(t,x) dt \le \int _{|t|< \sigma \delta \sqrt n} e^{-\frac{t^2}4} O\Big(\frac {t^2} n \Big)dt \le \frac Cn. $$
Turning to $I_3(t,x)$, we recall from Lemma \ref{lem: lambda}, that $b \in \B_\epsilon$, hence as a continous function on $\S$, it is bounded. Then
\begin{align*}
\abs{I_3(t,x)} ~\le~ \frac{t \normb{b}}{\sigma \sqrt{n}} e^{-\frac{1}{2}t^2} \abs{\exp\big( n f ( \frac{t}{\sigma \sqrt{n}}) \big) -1 } ~\le~ \frac{t \normb{b}}{\sigma \sqrt{n}} e^{-\frac{1}{4}t^2} \left( \frac{m_3 t^3}{6 \sigma^3 \sqrt{n}} + o(t^3/\sqrt{n}) \right),
\end{align*}
where we used again inequality \eqref{eq:feller}. Hence $\int \abs{I_3(t,x)}/t \, dt = O(1/n)$.
The integral over $I_4(t,x)$ is bounded independently of $x$ and vanishes at an exponential rate in $n$ since $\|N(t)\|\le c(1-\eta)^n$ by Corollary \ref{cor:Rt}.
}
\end{proof}
\section{The Bahadur-Rao Theorem for Products of Random Matrices}\label{sect:BahadurRao}
Now we are ready to prove our main result simultaneously for all types of matrices. We extend the approach for the one-dimensional case in \cite[Theorem 3.7.4]{Dembo1998}. Recall the definition $\Lambda(s)=\log k(s)$, such that $ \Lambda'(s) =\E_{\Q^s} S_1=: q$ and the Fenchel-Legendre transform of $\Lambda$ is given by$ \Lambda^*(q) = sq - \Lambda(s)$.
\begin{thm}
\label{thm:BahadurRao2}
Assume that $\mu$ satisfies $\condC$ and is non-arithmetic; or that \textit{(i-p)}~or \textit{(id)}~hold.
Let $q=\E_{\QQ^s} S_1 = \frac{k'(s)}{k(s)} $ for some $s \in \interior{I_\mu}$ and assume there is $0 < \epsilon < 1$ such that
\eqref{eq:iotamoment} holds.
\begin{enumerate}
\item Then
\begin{equation}
\label{eq:br upper bound}
\limsup_{n \to \infty} \, \sup_{x \in \S} \, \sup_{d \in [0, \infty)} \, e^{sd} \frac{\sqrt{n}e^{snq}}{k(s)^n} {\color{black} \Q_x}({S_n \ge nq+d}) < \infty.
\end{equation}
\item Consequently, there is $C < \infty$ s.t.~ for all $n \in \N$ and thereupon for each $u \ge nq$
\begin{equation}
\label{eq:br1}
{\color{black}\QQP[x]{S_n > u} }\le \frac {C k(s)^n}{\sqrt n e^{s u}}.
\end{equation}
\item For each fixed $\theta \ge 0$ it holds that
\begin{equation}
\label{eq:br uniform convergence}
\lim_{n \to \infty} \, \sup_{x \in \S} \sup_{d \in [0, \theta \sqrt{n})} \, \abs{ s \sigma \sqrt{2 \pi n} \, \frac{e^{s(nq +d)}}{k(s)^n} \, e^{\frac{d^2}{2 \sigma^2 n}} \E_{\Q_x}\Big[ \es(X_n) \1[{\{ S_n \ge nq+d\}}] \Big]- \es(x)} = 0.
\end{equation}
\item In particular, for all $x \in \S$,
\begin{equation}
\lim_{n \to \infty} \, s \sigma \sqrt{2 \pi n} \,e^{n \Lambda^*(q)} \E_{\Q_x}\Big[ \es(X_n) \1[{\{ S_n \ge nq\}}] \Big] = \es(x).
\end{equation}
\end{enumerate}
\end{thm}
Note that, using just the Chebyshev inequality and the definition of $k(s)$, one obtains in \eqref{eq:br1} the weaker upper bound
$$ {\color{black} \QQP[x]{S_n > u}} \le \frac{C k(s)^n}{e^{su}},$$
where the factor $1/\sqrt{n}$ does not appear.
\begin{proof}
All the results will be consequences of a general argument. Fix $\theta \ge 0$, but let $d \ge 0$ be arbitrary for the time being. Introduce $\Psi_n := s \sigma \sqrt{n}$ and
$$ J_n^d ~:=~ s \sigma \sqrt{2 \pi n} \frac{e^{snq} e^{sd}}{k(s)^n}~=~ \Psi_n \, \sqrt{2 \pi} \, e^{n \Lambda^*(q)} \, e^{sd}$$
as well as the normalized quantity
$$ W_n ~:=~ \frac{S_n - nq}{\sqrt{n}\sigma}, $$
then $\Prob_{{\QQ_x^s}}(W_n \le t ) = F_{n,x}(t)$. We obtain that
\begin{align*}
\frac{1}{\es(x)} \E_{\Q_x} \left({\es(X_n) \, \1[{\{S_n \ge nq + d \}}]}\right) ~&=~ \Erw[{{\QQ_x^s}}]{e^{n\Lambda(s)-sS_n} \, \1[{\{S_n \ge nq + d \}}]} \\
&=~ e^{- n \Lambda^*(q)} \Erw[{{\QQ_x^s}}]{e^{-s(S_n - nq)} \, \1[{\{S_n - nq\ge d \}}]} \\
&=~ e^{- n \Lambda^*(q)} \Erw[{{\QQ_x^s}}]{e^{-\Psi_n W_n} \, \1[{\{W_n \ge \frac{d}{\sqrt{n} \sigma} \}}]}
\end{align*}
Using the definition of $J_n^d$, we obtain
\begin{align*}
&~ J_n^d \frac{1}{\es(x)} \E_{\Q_x} \left({\es(X_n) \, \1[{\{S_n \ge nq + d \}}]}\right) \\
=&~ \sqrt{2 \pi} \Psi_n \, e^{sd}\, \int_{\frac{d}{\sigma \sqrt{n}}}^{\infty} e^{-\Psi_n t} \, dF_{n,x}(t) \\
=&~ \left. \sqrt{2 \pi} e^{sd} \Psi_n e^{-\Psi_n t} \, F_{n,x}(t) \right|_{\frac{sd}{\Psi_n}}^{\infty} + \sqrt{2 \pi} \, e^{sd}\, \int_{\frac{sd}{\Psi_n}}^{\infty} \Psi_n^2 e^{-\Psi_n t} \, F_{n,x}(t) \, dt\\
=&~ - \sqrt{2 \pi} \Psi_n \, F_{n,x}\left(\frac{sd}{\Psi_n}\right) + \sqrt{2 \pi} \, e^{sd}\, \int_{sd}^{\infty} \Psi_n e^{- t} \, F_{n,x}\left(\frac{t}{\Psi_n} \right) \, dt\\
=&~ \sqrt{2 \pi} \, e^{sd}\, \int_{sd}^{\infty} \Psi_n e^{- t} \, \left[ F_{n,x}\left(\frac{t}{\Psi_n} \right) - F_{n,x}\left(\frac{sd}{\Psi_n}\right)\right]\, dt\\
\end{align*}
Defining $h(t) := (1-t^2)\phi(t)$ and setting as before
$$ {\color{black} G_n(t) ~:=~ \Phi(t) + \frac{m_3}{\sigma^3 \sqrt{n}}(1-t^2)\phi(t) - \frac{b(x)}{\sigma \sqrt{n}} \phi(t) ~=~ \Phi(t) + \frac{m_3}{\sigma^3 \sqrt{n}}h(t) - \frac{b(x)}{\sigma \sqrt{n}} \phi(t) },$$
we want to use the Edgeworth expansion from Theorem \ref{thm:edgeworth} in order to calculate the asymptotics. Therefore,
\begin{align*}
&~ J_n^d \frac{1}{\es(x)} \E_{\Q_x} \left({\es(X_n) \, \1[{\{S_n \ge nq + d \}}]}\right) \\
=&~ \sqrt{2 \pi} \, e^{sd}\, \int_{sd}^{\infty} e^{- t} \, s \sigma \sqrt{n} \, \left( \left[ F_{n,x}\left(\frac{t}{\Psi_n} \right) - G\left(\frac{t}{\Psi_n} \right) \right] - \left[ F_{n,x}\left(\frac{sd}{\Psi_n}\right) - G\left(\frac{sd}{\Psi_n}\right) \right] \right)\, dt &~(=: I_1^d(n,x))\\
&~+ \sqrt{2 \pi} \, e^{sd}\, \int_{sd}^{\infty} \Psi_n e^{- t} \, \left[ \Phi\left(\frac{t}{\Psi_n} \right) - \Phi\left(\frac{sd}{\Psi_n}\right)\right]\, dt &~(=: I_2^d(n)) \\
&~+ \frac{m_3\sqrt{2 \pi}}{\sigma^3 \sqrt{n}} \, e^{sd}\, \int_{sd}^{\infty} \Psi_n e^{- t} \, \left[ h\left(\frac{t}{\Psi_n} \right) - h\left(\frac{sd}{\Psi_n}\right)\right]\, dt &~(=: I_3^d(n)) \\
&~- \frac{b(x)\sqrt{2 \pi}}{\sigma \sqrt{n}} \, e^{sd}\, \int_{sd}^{\infty} \Psi_n e^{- t} \, \left[ \phi\left(\frac{t}{\Psi_n} \right) - \phi\left(\frac{sd}{\Psi_n}\right)\right]\, dt &~(=: I_4^d(n)) \\
\end{align*}
\medskip
It follows from Theorem \ref{thm:edgeworth} that $$ \lim_{n \to \infty} \sup_{x \in \S} \, \sup_{d \ge 0} \, \abs{I_1^d(n,x)} = 0.$$
Using mainly that $\phi$ and $h$ have bounded derivatives, we are going to show that as well
\begin{equation}\label{eq:estimateI3} \lim_{n \to \infty} \sup_{d \ge 0} \abs{I_3^d(n)} = \lim_{n \to \infty} \sup_{d \ge 0} \abs{I_4^d(n)} =0. \end{equation}
Finally, considering $I_2^d(n)$, we are going to obtain two different estimates, namely
\begin{equation} \label{eq:estimateI2}\abs{I_2^d(n)} \le e^{-\frac{d^2}{2 \sigma^2 n}} \le 1, \end{equation}
and the refined estimate
\begin{equation} \lim_{n \to \infty} \sup_{d \in [0, \theta \sqrt{n}]} \abs{e^{\frac{d^2}{2 \sigma^2 n}} I_2^d(n) - 1} = 0. \label{eq:goodestimateI2}\end{equation}
Using the estimate \eqref{eq:estimateI2} allows to infer the upper bound \eqref{eq:br upper bound}, while the convergence result \eqref{eq:br uniform convergence} follows by using estimate \ref{eq:goodestimateI2}.
So it remains to prove Eqs. \eqref{eq:estimateI3} -- \eqref{eq:goodestimateI2}.
\medskip
\Step[2]: We consider $I_3^d$ and omit $I_4^d$, which can be treated along similar lines. A simple calculation shows that $h$ has a continuous derivative $h'$ with $\sup_{x \in \R} \abs{h'(x)} =: M < \infty$.
We compute
\begin{align*}
\abs{I_3^d(n)} ~=&~ \abs{ \frac{m_3\sqrt{2 \pi}}{\sigma^3 \sqrt{n}} \, e^{sd} \, \int_{sd}^{\infty} e^{-t} \, \left( \int_{sd/\Psi_n}^{t/\Psi_n} h'(r) \, \Psi_n \, dr \right) \, dt} \\
~=&~ \abs{ \frac{m_3\sqrt{2 \pi}}{\sigma^3 \sqrt{n}} \, e^{sd} \, \int_{sd}^{\infty} e^{-t} \, \left( \int_{sd}^{t} h'\left( \frac{r}{\Psi_n}\right) \, dr \right) \, dt}
=~ \abs{ \frac{m_3\sqrt{2 \pi}}{\sigma^3 \sqrt{n}} \, e^{sd} \, \int_{sd}^{\infty} \, h'\left( \frac{r}{\Psi_n}\right) \, \int_{r}^{\infty} e^{-t} \, dt \, dr } \\
\le&~ \frac{m_3\sqrt{2 \pi}}{\sigma^3 \sqrt{n}} \, e^{sd} \, \int_{sd}^{\infty} \, \abs{h'\left( \frac{r}{\Psi_n}\right)} e^{-r} \, dr
~\le~ \frac{m_3\sqrt{2 \pi}}{\sigma^3 \sqrt{n}} \, e^{sd} \, \int_{sd}^{\infty} \, M e^{-r} \, dr = M\, \frac{m_3\sqrt{2 \pi}}{\sigma^3 \sqrt{n}}
\end{align*}
\Step[3]: We are going to prove \eqref{eq:estimateI2} and \eqref{eq:goodestimateI2}. Therefore,
\begin{align}
I_2^d(n) ~=&~ e^{sd} \, \int_{sd}^\infty \Psi_n e^{-t} \, \left( \int_{sd/\Psi_n}^{t/\Psi_n} \, e^{-r^2/2} \, dr\right) \, dt ~=~ e^{sd} \, \int_{sd/\Psi_n}^\infty \Psi_n e^{-r^2/2} \, \left( \int_{\Psi_n r}^{\infty} \, e^{-t} \, dt\right) \, dr \\
=&~ e^{sd} \, \int_{sd/\Psi_n}^\infty \Psi_n e^{-r^2/2} \, e^{- \Psi_n r} \, dr ~=~ e^{sd} \,\left[ \left. - e^{- \Psi_n r - r^2/2} \right|_{sd/\Psi_n}^\infty - \int_{sd/\Psi_n}^\infty r e^{-r^2/2 - \Psi_n r} \, dr \right] \\
=&~ e^{-\frac{d^2}{2 \sigma^2 n}} - \, \int_{sd/\Psi_n}^\infty r \, e^{sd - \Psi_n r} \, e^{-r^2/2 } \, dr \label{eq:I2est}
\end{align}
For all $d \ge 0$, we have
$$ 0 \le \int_{sd/\Psi_n}^\infty r \, e^{sd - \Psi_n r} \, e^{-r^2/2 } \, dr \le \int_{sd/\Psi_n}^\infty r \, e^{-r^2/2 } \, dr ~=~ e^{-\frac{d^2}{2 \sigma^2 n}}$$
and thus \eqref{eq:estimateI2} follows.
\Step[4]: In order to prove \eqref{eq:goodestimateI2}, let $\epsilon > 0$ be arbitrary and choose $\delta$ such that $\delta + \frac{\theta \delta}{\sigma^2} + \frac{\delta^2}{\sigma^2} < \epsilon$. We separate the last integral in Eq. \eqref{eq:I2est} into
$$ \int_{sd/\Psi_n}^{sd/\Psi_n + \delta/\sigma} r \, e^{sd - \Psi_n r} \, e^{-r^2/2 } \, dr + \int_{sd/\Psi_n + \delta/\sigma}^\infty r \, e^{sd - \Psi_n r} \, e^{-r^2/2 } \, dr ~=:~ A(n) + B(n)$$
and see that by the restriction $d \le \theta \sqrt{n}$, it holds that $sd / \Psi_n \le {\theta}/{\sigma}$ and thus
$$ A(n) \le \frac{\delta}{\sigma} \frac{\theta + \delta}{\sigma} e^{-\frac{d^2}{2 \sigma^2 n}}.$$
Finally,
$$ B(n) \le e^{-s \delta \sqrt{n}} \int_{sd/\Psi_n + \delta/\sigma}^\infty r \, \, e^{-r^2/2 } \, dr = e^{-s \delta \sqrt{n}} \, e^{-(sd/\Psi_n + \delta/\sigma)^2/2 } \le e^{-s \delta \sqrt{n}} \, e^{-\frac{d^2}{2 \sigma^2 n}} .$$
Upon choosing $n_0$ such that $e^{- s \delta \sqrt{n}} \le \delta$ for all $n \ge n_0$, we obtain that for all $n \ge n_0$, $0 \le A(n) + B(n) \le \epsilon\, e^{-\frac{d^2}{2 \sigma^2 n}}$.
Thus we have proven that for all $\epsilon > 0$, there is $n_0$ such that for all $n \ge n_0$,
$$ \abs{e^{\frac{d^2}{2 \sigma^2 n}}I_2^d(n) - 1 } \le \epsilon.$$
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:BahadurRao}]
The main result is given above, while the formulas for $\sigma^2$ follow from Lemma \ref{lem: lambda} and Corollary \ref{cor:k}.
\end{proof}
\section{Tails of Stationary Solutions of Random Difference Equations}\label{sect:rde}
This section is devoted to Theorem \ref{thm:rde}. We start by giving an example for a matrix recursion from financial time series.
\begin{exa}
Consider the ARCH(2) process $Y_n$ defined by
$$ Y_n ~=~ \sigma_n \epsilon_n, \qquad \sigma_n^2 ~=~ a_1 Y_{n-1}^2 + a_2 Y_{n-2}^2 + 1, $$
where $\epsilon_n$ are i.i.d.~standard normal distributed random variables and $a_1, a_2 >0$ with $a_1 + a_2 < 1$.
\begin{enumerate}
\item Considering the squared process $(Y_n^2)_n$, we obtain a matrix recursion:
\begin{equation}
\left( \begin{array}{c}
\sigma_n^2 \\ Y_{n-1}^2
\end{array} \right) ~=~
\left( \begin{array}{cc}
a_1 \epsilon_{n-1}^2 & a_2 \\
\epsilon_{n-1}^2 & 0
\end{array} \right) \
\left( \begin{array}{c}
\sigma_{n-1}^2 \\ Y_{n-2}^2
\end{array} \right) +
\left( \begin{array}{c}
1 \\ 0
\end{array} \right) ~=:~ {\color{black} \mM_n} \vec{Y_{n-1}} + B_n,
\end{equation}
{\color{black} where $(\mM_n, B_n)_{n \in \N}$ is an i.i.d.~sequence in $\Mset \times \Rdnn$.}
{\color{black} The matrix $\mA_1 := \mM_1^\top$ satisfies condition $\condC$ (it suffices to assume that allowable matrices have full measure), and the moment condition \eqref{eq:iotamoment} is readily checked, since
\begin{align*}
\iota(\mA_1)^2 ~=~ \min_{x \in \Sp} \norm{\mA_1 x}^2 ~=~ \min_{\begin{subarray}{c}{x_1^2+x_2^2=1} \\ x_1, x_2 \ge 0 \end{subarray}} \, \left( a_1 \epsilon_0^2 x_1 + \epsilon_0^2 x_2 \right)^2 + a_2^2 x_1^2 ~\ge~ \big(\min\{a_1 \epsilon_0^2, a_2\} \big)^2.
\end{align*}
Consequently,
$\iota(\mA_1) \ge \min\{a_1 \epsilon_0^2, a_2 \}$, which has all negative moments up to order 1/2 since $\epsilon_1$ is a standard normal random variable.}
Kesten \cite[Theorem 3]{Kesten1973} gives the following sufficient condition for the existence of $\alpha >0$ such that $k(\alpha)=1$: There is $s_0 >0$ such that
$$ \E{ \left[ \big( \min_{i} \ \sum_{j} {\color{black} (\mM_1)_{i,j}} \big)^{s_0} \right]} ~\ge~ d^{s_0/2},$$
where $d$ is the dimension of the matrix {\color{black} (this is not a misprint, Kesten's condition is stated in terms of the matrix $\mM_1$)}. In our case, we have the estimate
$$ {\color{black} \E{ \left[ \big( \min_{i} \ \sum_{j} (\mM_1)_{i,j} \big)^s \right]} ~=~ \E \left[ \big( \min\{ a_1 \epsilon_0^2 + a_2, \epsilon_0^2\} \big)^s \right] ~\ge~ a_1^s \E (\epsilon_0^{2s}) }. $$
Since $\epsilon_1$ is unbounded, the right hand side tends to infinity as $s$ grows, thus there is $\alpha >0$ with $k(\alpha) >0$. Finally, the non-arithmeticity assumption holds since $\epsilon$ has a continuous distribution, and eigenvalues depend continuously on the entries of a matrix.
\item Kl\"uppelberg and Pergamenchtchikov showed in \cite[Lemma 2.7]{Klueppelberg2004}, that the process $(Y_n)_n$ has the same distribution as the process $(X_n)_n$ (if started with the same initial value), given by
$$ X_n ~=~ a_1 \eta_{1,n} X_{n-1} + a_2 \eta_{2,n} X_{n-2} + \eta_{3,n},$$
where $(\eta_{i,n})_n$ are independent sequences of i.i.d~standard normal random variables.
This leads to the matrix recursion
\begin{equation}
\left( \begin{array}{c}
X_n \\ X_{n-1}
\end{array} \right) ~=~
\left( \begin{array}{cc}
a_1 \eta_{1,n} & a_2 \eta_{2,n} \\
1 & 0
\end{array} \right) \
\left( \begin{array}{c}
X_{n-1} \\ X_{n-2}
\end{array} \right) +
\left( \begin{array}{c}
\eta_{3,n} \\ 0
\end{array} \right) ~=:~ {\color{black} \mM_n} \vec{X_{n-1}} + B_n
\end{equation}
It can be shown that these matrices satisfy assumptions \textit{(i-p)}~as well as \textit{(id)}, and it is proved in \cite[Lemma 3.2]{Klueppelberg2004} that there exists $\alpha > 0$ with $k(\alpha)=1$.
\end{enumerate}
\end{exa}
{\color{black}
Further instances of the equation $\R \eqdist \mM R + B$, with matrices satisfying the assumptions of Theorem \ref{thm:rde}, appear e.g.~ in \cite{Basrak2002} (GARCH-processes, nonnegative matrices), \cite{Wang2013} (multitype branching processes with immigration in random environment, nonnegative matrices), \cite{Behme2012} (stationary solutions of multivariate generalized Ornstein-Uhlenbeck processes, invertible matrices), to name just a few.
An extension of the methods used below applies to provide exact tail asymptotics for random variables $R$ which are fixed points of \emph{multivariate smoothing transforms}, i.e.
satisfying
\begin{equation}\label{eq:SFPE} R \eqdist \sum_{i=1}^N \mM_i R_i + B,\end{equation}
where $N \ge 2$ is a fixed integer, $R$ and $R_i$ are i.i.d.~and independent of the random matrices $\mA_i$ and the random vector $B$. The details are worked out in \cite{BM2015}.
Heavy tail properties of such $R$ were studied in \cite{BDG2011,BDMM2013,Mirek2013} and a result similar to \eqref{eq:RDE} was obtained, there $\alpha=\max\{s > 0 : \k(s) = 1/N\}$, but only in the first reference, which studies matrices satisfying $\condC$, it could be shown that $K >0$, in the latter two references, only partial results were obtained.}
{\color{black} \subsection{Outlining the proof of Theorem \ref{thm:rde}}
Now we explain how the proof of Theorem \ref{thm:rde} is given by a sequence of lemmata, the proofs of which are {\color{black} quite} technical and therefore postponed to the subsequent section, for a better stream of arguments. First, we have to introduce some notation.
{\color{black}
\begin{notation}
Given a random element $(\mM, B) \in M(d \times d, \R) \times \R^d$, let $(\mM_n, B_n)_{n \in \N}$ be a sequence of i.i.d.~copies of $(\mM, B)$, defined on a probability space $(\Omega, \F, \Prob)$. Let $X_0 : \Omega \to \Sd$ be a random variable and $(\Prob_x)_{x \in \Sd}$ be a family of probability measures on $\Omega$, such that $(\mM_n, B_n)_{n \in \N}$ have the same law as under $\Prob$, while $\P[x]{X_0=x}=1$.
Write ${\blue\matrix{\daleth}}_n:= \mM_n^\top \cdots \mM^\top_1$, $X_n^*:={\blue\matrix{\daleth}}_n \as X_0$, $S_n^* := \log \abs{{\blue\matrix{\daleth}}_n X_0}$.
\end{notation}
As mentioned before, we will apply the results obtained in the previous sections to the matrix $\mA_1 := \mM_1^\top$. Writing $\mu$ for the law of $\mA_1=\mM_1^\top$ under $\Prob$, and defining the measures $\Q_x$ as in Subsection \ref{sect:defQ}, we have the following identities, valid for all $x \in \S$:
\begin{align}
\Q_x \big( (X_0, (\mA_n)_{n \in \N}) \in \cdot \big) ~=~ \P[x]{ (X_0, (\mM_n^\top)_{n \in \N}) \in \cdot)} \\
\Q_x \big( (X_n, S_n)_{n \in \N} \in \cdot \big) ~=~ \P[x]{ (X_n^*, S_n^*)_{n \in \N}) \in \cdot)}.
\end{align}
If not explicitly stated otherwise, all appearing quantities below will be defined in terms of the sequences $(\mA_n)_{n \in \N}$, for example $k(s)$.
From now on, we fix $\alpha >0$ such that $k(\alpha)=1$ and set $q:=\E_{\Q^s}^\alpha S_1$. We will assume throughout that the assumptions of Theorem \ref{thm:rde} are in force.
}
Using the identifications from above, the SLLN in Proposition \ref{prop:FK} yields that
$$ \lim_{n \to \infty} \frac{1}{n} \log \norm{{\blue\matrix{\daleth}}_n} ~=~k'(0) <0 \qquad \Pfs, $$
which allows to infer that there is a unique solution (in distribution) to the equation $R \eqdist \mM R +B$, see e.g. \cite[Theorem 1.1]{Bougerol1992}.
\medskip
{\color{black}
The fundamental idea is to compare the behavior of $\skalar{x,R}$ with that of $\abs{{\blue\matrix{\daleth}} x} $. Therefore, we use that for $R_k$ being i.i.d.~copies of $R$ and independent of $(\mM_k, B_k)$, we have that for all $n \in \N$,
\begin{eqnarray*}
R&\eqdist& \mM_1 R_1+B_1 \eqdist \mM_1 \cdots \mM_n R_n + \sum_{k\le n} \mM_1 \cdots \mM_{k-1}B_k.
\end{eqnarray*}
Consequently, for any $x \in \Sd$,
\begin{eqnarray*}
\skalar{x,R} ~\eqdist~ \skalar{\color{black}{x,\mM_1} \cdots \mM_n R_n} + \sum_{k\le n} \skalar{x,\mM_1 \cdots \mM_{k-1}B_k} ~\ge~ \skalar{{\blue\matrix{\daleth}}_n x, R_n} - \sum_{k\le n} \abs{\skalar{{\blue\matrix{\daleth}}_{k-1} x,B_k}}
\end{eqnarray*}
We are going to consider sets where first term dominates, while the sum is comparably small. In order to estimate the scalar product $\skalar{{\blue\matrix{\daleth}}_n x, R_n}$ from below by $\abs{{\blue\matrix{\daleth}}_n x}$, we will use the following lemma:
\begin{lem}\label{lem:cones}
Let the assumptions of Theorem \ref{thm:rde} hold.
Then for all $D >0$ there are $J < \infty$, $\kappa_j>0$ and $c_j>0$, $1 \le j \le J$, and disjoint subsets $\S_j \subset \Sd$, such that
\begin{equation}\label{eq:cones}\P{ \frac{R}{\abs{R}} \in \S_j \ \text{ and } |R|> \frac{D}{c_j}} ~\ge~ \kappa_j \end{equation}
and moreover
\begin{equation}
\label{eq: 4.2cones}
\Rd \subset \bigcup_{j=1}^J \S_j^*,
\end{equation}
where $\S_j^*$ are the cones
$$ \S_j^* := \{ y \in \Rd \, : \, \skalar{y,x} \ge c_j \abs{y} \text{ for all $x \in \S_j$}\}. $$
If $\mu$ satisfies (C), then the same statement is valid, but for $\S_j$ being disjoint subsets of $\Sp$ and with \eqref{eq: 4.2cones} replaced by
$$ \R_{\ge} \subset \bigcup_{j=1}^J \S_j^*.
$$
\end{lem}
The proof of the lemma will be given in Section \ref{sect:rde proofs}.
The lemma now allows for the following comparison: If $R_n \in \S_j$ and ${\blue\matrix{\daleth}}_n x \in \S_j^*$, it follows that $\skalar{{\blue\matrix{\daleth}}_n x, R_n} \ge c_j \abs{{\blue\matrix{\daleth}}_n x} \abs{R_n}$.
}}
\medskip
{\color{black} As the next step, we use this comparison in more detail. }
Given constants $C_0, \delta >0$ (which will be chosen later), let $D = e^{C_0} \sum_{k=0}^\infty e^{-k \delta} = e^{C_0}/(1-e^{-\delta})$. Let $t \ge 0$ and define $n_t = \lceil \log t/q \rceil$.
\begin{eqnarray}
V_{n,t} &=& \Big\{ \color{black}{S_n^*} \ge n_t q \ \mbox{ and } \ \log |B_{k+1}| + S_k^* \le n_t q + C_0 -(n-k)\delta \ \forall k< n \Big\} \label{v1}\\
V_{n,t}^j&=& V_{n,t} \cap \big\{ {\blue\matrix{\daleth}}_n X_0\in \S_j^* \big\}\cap \big\{ R_n\in \S_j \ \mbox{ and } |R_n|> 2 \frac{D}{c_j} \big\}, \label{v2}\\
\widetilde V_{n,t} &=& \bigcup_j V_{n,t}^j. \label{v3}
\end{eqnarray}
Then we have the following lemma, the short proof of which we give immediately.
\begin{lem}
\label{lem:pos:12.3} For all $t \ge 0$,
$$
\P{\skalar{x,R}>Dt} \ge \P[x]{\bigcup_n \widetilde V_{n,t}}.
$$ Moreover, for all $n \in \N$,
$$\P[x]{\widetilde V_{n,t}}\ge (\min_j\, \kappa_j) \,\P[x]{V_{n,t}} =: \kappa_0 \P[x]{V_{n,t}}.$$
\end{lem}
\begin{proof}
Recall that $\P[x]{X_0=x}=1$. Thus for every $n$ on the set $ V_{n,t}^j$ we have under $\Prob_x$
{\color{black} \begin{eqnarray*}
\bigg\langle \mM_1 \cdots \mM_{n} R_n + \sum_{k\le n}\mM_1 \cdots \mM_{k-1}B_k,x \bigg\rangle &\ge&
\big\langle R_n , {\blue\matrix{\daleth}}_n x \big\rangle - \sum_{k\le n}\big|\big\langle B_k, {\blue\matrix{\daleth}}_{k-1}x \big\rangle\big|\\
&\ge&
c_j |R_n||{\blue\matrix{\daleth}}_n x| - \sum_{k\le n}| B_k| |{\blue\matrix{\daleth}}_{k-1}x |\\
&\ge& 2D e^{n_t q} - e^{n_t q} e^{C_0} \sum_{k \le n} e^{-(n-k)\delta} \\
&\ge& Dt
\end{eqnarray*} }
To prove the second part of the Lemma we use the independence of ${\blue\matrix{\daleth}}_n$ and $R_n$, the disjointness of $\S_j$ and the fact that $R \eqdist R_n$ to deduce
$$
{\color{black} \P[x]{\widetilde V_{n,t}} = {\mathbb P}_x \bigg(\bigcup_j V^j_{n,t}\bigg) = \sum_j \P[x]{ V^j_{n,t}} \ge \kappa_0 \sum_j \P[x]{V_{n,t} \cap \big\{ {\blue\matrix{\daleth}}_n x\in \S_j^* \big\} } = \kappa_0 \P[x]{V_{n,t}}.}
$$
\end{proof}
The final burden will then be to prove the following Lemma:
\begin{lem}\label{lem:finallemma}
There is $\eta >0$ and $T_0 >0$ such that for all $t \ge T_0$,
$${\color{black} \P[x]{\bigcup_n \widetilde V_{n,t}} ~\ge~ \eta t^{-\alpha}.}$$
\end{lem}
In fact, in its proof, we will for each fixed $t$ only consider a subset $K_t \subset \color{black}{\N}$ of integers close to $n_t:= \lfloor \log t/q \rfloor$, with {\color{black} $q = \E_{\Q^\alpha} S_1 > 0$}. We will choose $K_t \subset [n_t - \sqrt{n_t}, n_t]$ and prove that readily ${\color{black} \P[x]{\bigcup_{k \in K_t} \widetilde V_{k,t}}} ~\ge~ \eta t^{-\alpha}$, using the inclusion--exclusion formula.
Therefore, we will use the following technical result, which finally fixes $C_0$ and $\delta$. It is here where the Bahadur-Rao theorem enters.
\begin{lem} \label{lem:estimate Vn}{\color{black} Assume that $\alpha \in \interior{I_\mu}$ and that
\begin{equation}\label{eq:moment conditions2}
\E \norm{\mA}^{\alpha+\epsilon} \iota(\mA)^{-\epsilon} < \infty, \qquad 0 < \E \abs{B}^{\alpha + \epsilon} < \infty.
\end{equation}}
Then
there are constants $\delta, C_0,D_1,D_2,N_0 >0$ such that {\color{black} for all $x \in \S$}
$$
D_1 \cdot \frac{k(\alpha)^n}{\sqrt{n_t} e^{\alpha n_t q}}
\le {\color{black} \P[x]{V_{n,t}}} \le
D_2 \cdot \frac{k(\alpha)^n}{\sqrt{n_t} e^{\alpha n_t q}}.
$$ for all $\lceil \log t / q \rceil = n_t > N_0$ and every $n_t-\sqrt{n_t} \le n \le n_t - \sqrt{n_t}/2$.
For the assertion of this lemma to hold, $k(\alpha) =1$ is not necessary, we only need that $k'(\alpha)>0$, then still $q:= \E_{\Q^\alpha} S_1$.
\end{lem}
Summing up what has been said before, we now are able to prove Theorem \ref{thm:rde}.
\begin{proof}[Proof of Theorem \ref{thm:rde}]
{\color{black} We already mentioned that the assumptions of Theorem \ref{thm:rde} guarantee the existence and uniqueness (in distribution) of a solution $R$ to $R \eqdist \mM R+B$, see e.g.~\cite[Theorem 1.1]{Bougerol1992}
The lower bound for the tail behavior then follows by combining Lemmas \ref{lem:pos:12.3} and \ref{lem:finallemma}.}
\end{proof}
\section{Proofs}\label{sect:rde proofs}
{\color{black}
\subsection{On the support of $R$}
As mentioned before, in order to prove Lemma \ref{lem:cones}, we have to show that $\supp R$ is unbounded in ``enough'' directions. To make this statement precise, introduce
the {\em asymptotic support} of $R$:
Consider the compactification $\overline{\Rd} := \Rd \cup \Sd_\infty$ of $\Rd$, with
$$ \Rd \ni x_n \text{ converges to } y \in \Sd_\infty = \Sd \quad \Leftrightarrow \quad \lim_{n \to \infty} \frac{x_n}{\abs{x_n}} = y \text{ and } \lim_{n \to \infty} \abs{x_n}=\infty $$
We will study the set
$$ V(R) := \{ y \in \Sd \, : \, \exists (r_n)_n \subset \supp \, R \, : \, \lim_{n \to \infty} \frac{r_n}{\abs{r_n}} = y \text{ and } \lim_{n \to \infty} \abs{\color{black}{r_n}}=\infty \}.$$
Using diagonal sequences, one obtains that the set $V(R)$ is indeed closed and thus, as a subset of $\S$, even compact.
An important result is that the set $V(R)$ is invariant under the action of $\Gamma^*=[\supp \mM]$ on the sphere: Let $y \in V(R)$, with associated sequence $r_n$. Then $\matrix{m} r_n +b \in \supp \, R$ for all $(\matrix{m},b) \in \supp(\mM,B)$, and still $\abs{\matrix{m} r_n + b} \to \infty$, with the ratio $\abs{\matrix{m} r_n +b}/\abs{\matrix{m} r_n}$ tending to one. Hence,
$$ \matrix{m} \as y = \lim_{n \to \infty} \left( \frac{\matrix{m} r_n}{\abs{ \matrix{m} r_n}} + \frac{b}{\abs{\matrix{m} r_n}} \right)= \lim_{n \to \infty} \frac{\matrix{m} r_n +b}{\abs{ \matrix{m} r_n + b}} , $$
and thus $\matrix{m} \as y \in V(R)$.
We have the following result about $V(R)$ for $\mA$ being nonnegative.
\begin{prop}\label{lem:supp nonnegative}
Under the assumptions of Theorem \ref{thm:rde}, let $\mA$ be nonnegative and satisfy $\condC$.
\begin{enumerate}
\item Assume that $B$ is nonnegative. Then $V(R) \cap \Sp \neq \emptyset$.
\item If $V(R) \cap \Sp \neq \emptyset$, then readily $V(R) \cap \interior{\Sp} \neq \emptyset$.
\end{enumerate}
\end{prop}
}
\begin{proof
\Step[1]:
If $B$ is nonnegative as well, then $\supp R \subset \Rdnn$. Moreover, since $B \neq 0$, there is nonzero $r \in \supp R$.
But then as well ${\color{black} R_n^r := } \mM_1 \cdots \mM_n r + \sum_{k \le n} \mM_1 \cdots \mM_{k-1} B_k \in \supp R$, in particular,
$$ \abs{R_n^r} ~\ge~ \abs{\mM_1 \cdots \mM_n r} ~\eqdist~ \abs{\mM_n \cdots \mM_1 r}.$$
We assumed that $\mA = \mM^\top$ satisfies $\condC$; but $\condC$ holds for $\mA$ if and only if it holds for $\mA^\top$. Hence Proposition \ref{prop:FK} applies and gives under $ ^*\Q_r^\alpha$ (which denotes the measure constructed in the same manner as $\Q_r^\alpha$, but using the law of $\mA^\top=\mM$),
$$ \lim_{n \to \infty} \frac{S_n}{n} ~=~ \lim_{n \to \infty} \frac{1}{n} \log \abs{\mM_n \dots \mM_1 r} ~=~ k'(\alpha)/k(\alpha) >0.$$
But finite marginal distributions of $^*\Q_r^\alpha$ and $\Prob$ are equivalent, and thus we have that the sequence $(\abs{\mM_n \dots \mM_1 r})_n$ is unbounded, hence $ \supp\, R$ is unbounded as well.
\Step[2]:
As shown above, the set $V(R)$ is invariant under $\Gamma^*$. But by \cite[Lemma 4.3]{BDGM2014}, $V(\Gamma^*)$ is the unique closed minimal $\Gamma^*$-invariant subset, hence $V(R)$ contains $V(\Gamma^*)$. In particular, there is $y \in \interior{\Sp}$ with $y \in V(\Gamma^*) \subset V(R)$.
\end{proof}
{\color{black}
\begin{proof}[Proof of Lemma \ref{lem:cones}]
We consider seperately the three cases of matrices.
\Case[1]: Assume $\mA$ is nonnegative and satisfies $\condC$ and that $\supp R \cap \Rdnn$ is unbounded. Then, by Lemma \ref{lem:supp nonnegative}, there is $y_0 \in \interior{\Sp} \cap V(R)$. It holds that $\min_{x \in \Sp} \skalar{x,y_0} \ge \min_{1 \le i \le d} (y_0)_i >0$. Let $\delta >0$ such that $B_\delta(y_0) \cap \Sp \subset \interior{\Sp}$. Then there is $c >0$ such that $\min_{x \in \Sp} \min_{y \in B_\delta(y_0)} \skalar{x,y} \ge c$, and we can set $J:=1$ and $\S_1 := B_\delta(y_0)$.
\medskip
\Case[2]: Assume that $\mA \in GL(d, \R)$, satisfying $\textit{(id)}$. Due to the density assumption, which is invariant under taking the transpose, there is in particular a proximal matrix $\matrix{m} \in \Gamma^*$ with attracting eigenvector $v_\matrix{m}$. By \cite[Lemma 8.1]{AM2010}, the set $V(R)$ is non-empty, moreover, there are $y_1, y_2 \in V(R)$ such that
$\skalar{y_1, v_\matrix{m}} >0$ and $\skalar{y_2, (-v_\matrix{m})}>0$. Since $V(R)$ is $\Gamma^*$-invariant, it follows that $\matrix{m}^n \as y_1$ and $\matrix{m}^n \as y_2$ are in $V(R)$ for all $n \in \N$, hence $v_\matrix{m}$ and $-v_\matrix{m}$ are in $V(R)$.
Observe that the operator $\Pst[\alpha]$, defined in \eqref{eq:Pst}, leaves $\Cbf{V(R)}$ invariant. Moreover, due to the density assumption, the measure $\nu := (\Pst[\alpha])^{n_0} \frac{1}{2}( \delta_{v_\matrix{m}} + \delta_{-v_\matrix{m}})$ has a density with respect to the volume measure on $\Sd$, and in particular, gives mass zero to any hyperspace, and is still supported on (a subset of) $V(R)$ and is symmetric, i.e.~$\nu(A)=\nu(-A)$ for all $A \subset \Sd$. Then one can proceed as in \cite[Lemma 2.7 \& Lemma 2.8]{Guivarch2012}---these are the counterpart of Lemma \ref{lem:prop:qn} for invertible matrices---to show that $((\Pst[\alpha])^n) \nu (\1[V(R)]) \ge c \kappa(\alpha)^n =c$ for some $c>0$ and all $n \in \N$.
Together with the compactness of $V(R)$ this yields that, using Prokhorov's Theorem,
$$ \frac{1}{n} \sum_{k=0}^{n-1} (\Pst[\alpha])^k \nu $$
is a weakly compact sequence and therefore has a subsequential limit $\nust[\alpha]$, which is a probability measure on $V(R)$ that satisfies $\Pst[\alpha] \nust[\alpha] = \nust[\alpha]$.
By Prop. \ref{prop:transferoperators}, \eqref{suppnus} and \eqref{nust}, it holds for all $x \in \S$ that
$$ \min_{x \in \S} \int_{V(R)} \, \abs{\skalar{x,y}}^s \, \nust(dy) ~=~ \min_{x \in \S} \frac{1}{c} \es(x) ~:=~\epsilon >0.$$
This shows that for all $x \in \Sd$ there is $y \in V(R)$ such that $\skalar{x,y} \ge 2^{-1/s} \epsilon^{1/s}$ due to the symmetry of $\nust$. Hence we can find a partition of $V(R)$ into a finite number of sets (use compactness), such that the assertions of the Lemma hold.
\medskip
\Case[3]: Assume that $\mA \in GL(d,\R)$ satisfies $\textit{(i-p)}$, and that there is no proper closed convex cone, which is $\Gamma^*$-invariant. Then it is shown \cite[after Theorem 5.1]{Guivarch2012}, that $V(R)$ contains the pre-image of $V(\Gamma^*)$ under the projection $\Sd \to \mathbb{P}^{d-1}$ (this is called Case I there). Using that $\mM=\mA^\top$ satisfies \textit{(i-p)}~as well, if it is satisfied by $\mA$, we use \cite[Theorem 2.17]{Guivarch2012} to infer the existence of a symmetric probability measure $\nust[\alpha]$, which is supported in (a subset of) $V(R)$. Then we can conclude as above.
\end{proof} }
\subsection{Auxiliary Lemma}
Next we are going to prove Lemma \ref{lem:estimate Vn}.
\begin{proof}[Proof of Lemma \ref{lem:estimate Vn}]
\Step[1]: Denoting $U_{n,t} := \Big\{ S^*_n \ge n_t q \Big\}$ and
\begin{eqnarray*}
W_{j,{n,t}} &:=& \Big\{ S^*_j +\log|B_{j+1}|> n_t q + C_0 - (n-j)\delta \Big\},\\
\end{eqnarray*}
we have that
$$
{\color{black} \P[x]{V_{n,t}} = \P[x]{U_{n,t}} - {\mathbb P}_x \bigg({\bigcup_{j<n} (U_{n,t}\cap W_{j,{n,t}})}\bigg)}.
$$
Using the Bahadur-Rao type result \eqref{eq:br uniform convergence}, we estimate ${\color{black}\P[x]{U_{n,t}}}$ from below (with $d=q(n_t-n)$, $\theta = q+1$, $\vartheta_1 = \inf_{x,y} \frac{\est[\alpha](x)}{\est[\alpha](y)}$), namely, there is $N_0 \in \N$ and {\color{black}$\vartheta_2 >0$}, such that for all $n \ge N_0$, the following estimate holds:
\begin{equation}
\label{eq: pos:star}
{\color{black} \P[x]{U_{n,t}}} = \Q_x (S_n > n_t q) \ge \frac{\vartheta_1}{r_{\alpha}(x)} \E_{\Q_x} \Big[ \es[\alpha](X_n) {\bf 1}_{\{ S_n \ge nq + d \}} \Big] \ge \vartheta_2 \cdot
\frac{k(\alpha)^n}{\sqrt{n_t} e^{\alpha n_t q}}.
\end{equation}
Similarly \eqref{eq:br1} provides an upper estimate for $\P{U_{n,t}}$ which in particular proves the upper bound in the lemma.
Therefore it is sufficient to prove that
\begin{equation}\label{aim}
{\color{black} {\mathbb P}_x} \bigg( \bigcup_{j<n} (U_{n,t}\cap W_{j,{n,t}}) \bigg) \le \vartheta \cdot \frac{k(\alpha)^n}{\sqrt{n_t} e^{\alpha n_t q}}
\end{equation}
for some $\vartheta<\vartheta_2/2$. In fact, we are going to show that {\color{black}$\vartheta$} can be made arbitrarily small by choosing $C_0$ large.
\medskip
\Step[2]: Let us denote by $\mu_{\mA,B}$ the joint law of $(\mA,B)$.
By decomposing the sets $W_{j,{n,t}}$ further, depending on the overshoot of $S_j^* + \log \abs{B_{j+1}}$, we obtain \begin{align*}
&~{\color{black} {\mathbb P}_x} \bigg( \bigcup_{j<n} (U_{n,t}\cap W_{j,{n,t}}) \bigg)
~\le~ \sum_{j<n} {\color{black}\P[x]{ U_{n,t}\cap W_{j,{n,t}}}}\\
=&~ \sum_{j<n}\sum_{m\ge 0} {\mathbb P} \bigg( {
\frac{e^{n_t q + C_0 + m}}{e^{ (n-j)\delta}} \le |B_{j+1}| {\color{black} |{\blue\matrix{\daleth}}_j x|} < \frac{e^{n_t q + C_0 + m+1}}{e^{ (n-j)\delta}}\ \mbox{ and }\ \color{black}{ |{\blue\matrix{\daleth}}_n x|}> e^{n_tq}
} \bigg) \\
\le &~ \sum_{j<n}\sum_{m\ge 0} \int {\mathbb P}\bigg( {
\frac{e^{n_t q + C_0 + m}}{e^{ (n-j)\delta}} \le |b| |{\blue\matrix{\daleth}}_j x| < \frac{e^{n_t q + C_0 + m+1}}{e^{ (n-j)\delta}}\ \mbox{ and }\ \|{\blue\matrix{\daleth}}_{j+2}^n\|\|\ma^\top\||{\blue\matrix{\daleth}}_j x|> e^{n_tq}
} \bigg) \mu_{\mA,B}(d\ma,\, db) \\
\le&~ \sum_{j<n} \sum_{m\ge 0} \int
{\mathbb P} \bigg(
|{\blue\matrix{\daleth}}_j x| \ge \frac{ e^{n_t q + C_0 +m}}{ |b| e^{ (n-j)\delta}} \bigg)
\cdot {\mathbb P} \bigg( \|\mPi_{n-j-1}\|> \frac{|b|e^{ (n-j)\delta}}{\|\ma\| e^{C_0+ m + 1}}
\bigg) \mu_{\mA,B}(d\ma,\,db)\\
\end{align*}
To estimate further, we have to consider separately the cases where $\abs{b}$ is small resp. large, for we can apply the Bahadur-Rao estimate only in the first case.
More precisely, we split the integral into two integrals over the set
\begin{equation} \label{eq:defTheta} \Theta := {\{|b| \le e^{(n_t-j)q + C_0 - (n-j)\delta + m}\}} \end{equation}
and its complement $\Theta^c$, respectively.
\medskip
\Step[3]: In this step, we estimate
$$ I := \sum_{j<n} \sum_{m\ge 0} \int \1[\Theta](b) \,
{\mathbb P} \bigg(
|{\blue\matrix{\daleth}}_j x| \ge \frac{ e^{n_t q + C_0 +m}}{ |b| e^{ (n-j)\delta}} \bigg)
\cdot {\mathbb P} \bigg( \|\mPi_{n-j-1}\|> \frac{|b|e^{ (n-j)\delta}}{\|\ma\| e^{C_0+ m + 1}}
\bigg) \mu_{\mA,B}(d\ma,\,db).$$
{\color{black}
In this step, we also choose $\delta$. $C_0$ will be a free parameter until Step 5, and it is important to notice, that all appearing constants are independent of $C_0$.
On $\Theta$, $e^u:= \frac{ \exp({n_t q + C_0 +m})}{ |b| \exp({ (n-j)\delta}) } \ge \exp(jq)$, thus the estimate \eqref{eq:br1} applies to the first probability in $I$ and yields
\begin{equation}\label{eq:aa1} \P{ \abs{{\blue\matrix{\daleth}}_j x} \ge e^u } ~=~ \P[x]{S_n^* \ge u} ~=~ \QP[x]{S_n \ge u} ~\le~ \frac{C k(\alpha)^j}{\sqrt{j} \, e^{\alpha u}} ~=~ \frac{C k(\alpha)^j \abs{b}^\alpha e^{ \alpha (n-j) \delta}} {\sqrt{j}\, e^{\alpha (n_t q +C_0 +m)} }, \end{equation}
where $C$ is given by Theorem \ref{thm:BahadurRao2} and only depends on $\alpha$.
In order to estimate the second probability in $I$, we use the Markov inequality with the function $x \mapsto x^{\beta}$, where we choose $\beta >0$ such that $\beta < \alpha$ and $k(\beta) < k(\alpha)$. This is possible since $k'(\alpha)>0$. Then, by Corollary \ref{cor:ksEs},
$$ \E \norm{\mPi_{n-j-1}}^{\beta} ~\le~\frac{1}{c_\beta} k(\beta)^{n-j-1} ~=~ \frac{1}{c_\beta k(\beta)} k(\alpha)^{n-j} \left( \frac{k(\beta)}{k(\alpha)} \right) ^{n-j}$$
with $c_\beta >0$ given by Prop. \ref{lem:prop:qn}.
We obtain, applying the Markov inequality as described above,
\begin{align}
\label{eq:aa2} {\mathbb P} \bigg( \|\mPi_{n-j-1}\|> \frac{|b|e^{ (n-j)\delta}}{\|\ma\| e^{C_0+ m + 1}}
\bigg) ~ &\le ~ \frac{\E \big[\norm{\mPi_{n-j-1}}^\beta \big] \norm{\ma}^\beta e^{\beta(C_0 + m + 1)} }{\abs{b}^\beta e^{\beta(n-j)\delta}}\\ ~&\le~ \frac{k(\alpha)^{n-j} \norm{\ma}^\beta e^{\beta(C_0 + m + 1)} }{k(\beta) c_\beta \abs{b}^\beta e^{\beta(n-j)\delta}} \left( \frac{k(\beta)}{k(\alpha)} \right) ^{n-j} \nonumber
\end{align}
Define $\xi:= \alpha - \beta>0$. Using the estimates \eqref{eq:aa1} and \eqref{eq:aa2} in $I$ and simplifying terms, we obtain
\begin{equation}
\label{eq:aa3} I ~ \le ~ \sum_{j<n} \sum_{m\ge 0} \int \1[\Theta](b) \, \frac{C e^\beta}{c_\beta k(\beta )} \frac{1}{\sqrt{j}} \frac{k(\alpha)^n \abs{b}^\xi \norm{\ma}^\beta e^{\xi(n-j)\delta}} {e^{\alpha n_t q} \, e^{\xi(C_0 +m)} } \, \left( \frac{k(\beta)}{k(\alpha)} \right) ^{n-j} \, \mu_{\mA,B}(d\ma,\,db)
\end{equation}
We estimate further by omitting the indicator $\1[\Theta](b)$, integrating, setting $C' := C e^\beta /c_\beta k(\beta)$ and using Fubini's theorem:
\begin{align}
\label{eq:aa4} I ~ \le& ~ \sum_{j<n} \sum_{m\ge 0} C' \frac{1}{\sqrt{j}} \frac{k(\alpha)^n e^{\xi(n-j)\delta}} {e^{\alpha n_t q} \, e^{\xi(C_0 +m)} } \, \left( \frac{k(\beta)}{k(\alpha)} \right) ^{n-j} \, \E \big( \norm{\mA}^\beta \abs{B}^\xi \big) \\
\nonumber =&~ \frac{C' k(\alpha)^n }{e^{\xi C_0} e^{\alpha n_t q}} \left( \sum_{m \ge 0} e^{-\xi m} \right) \left( \sum_{j<n} \frac{1}{\sqrt{j}} \left( \frac{k(\beta) e^{\xi \delta}}{k(\alpha)} \right) ^{n-j} \right) \, \E \big( \norm{\mA}^\beta \abs{B}^\xi \big)
\end{align}
Recall that we chose $\beta <\alpha$ such that $k(\beta) < k(\alpha)$. Thus, we can choose a (small) $\delta >0$, such that
\begin{equation} \label{eq:exi} e^{\xi \delta} k(\beta) < k(\alpha), \end{equation} and apply Lemma \ref{lem:sumineq} (see below) with $\rho=e^{\xi \delta} k(\beta) /k(\alpha)$ to infer that the sum over $j$ is bounded by a constant times $1/\sqrt{n}$.
On $\E( \norm{\mA}^\beta \abs{B}^\xi)$ we can apply H\"older's inequality with $p_1=\alpha/(\alpha-\xi)=\alpha/\beta$ and $p_2= \alpha/ \xi$ to infer
$$ I ~\le~ \frac{C' k(\alpha)^n }{e^{\xi C_0} e^{\alpha n_t q}} \, \frac{1}{1-e^{-\xi}} \, \frac{D_3}{\sqrt{n}} \, \E \norm{\mA}^\alpha \, \E \abs{B}^\alpha ~=:~ \frac{C'' k(\alpha)^n }{\sqrt{n} e^{\xi C_0} e^{\alpha n_t q}} ~\le~ \frac{C'''}{e^{\xi C_0}} \cdot \frac{ k(\alpha)^n }{\sqrt{n_t} e^{\alpha n_t q}} $$
for a finite constant $C'''$, which does not depend on $n$ or $t$ or $C_0$. Note that we were allowed to replace $n$ by $n_t$ in the final expression, since $n_t-\sqrt{n_t} \le n \le n_t - \sqrt{n_t}/2$ by assumption.
}
\medskip
\Step[4]:
Now, to estimate
$$II~:=~\sum_{j<n} \sum_{m\ge 0} \int \1[\Theta^c](b) \,
{\mathbb P} \bigg(
|{\blue\matrix{\daleth}}_j x| \ge \frac{ e^{n_t q + C_0 +m}}{ |b| e^{ (n-j)\delta}} \bigg)
\cdot {\mathbb P} \bigg( \|\mPi_{n-j-1}\|> \frac{|b|e^{ (n-j)\delta}}{\|\ma\| e^{C_0+ m + 1}}
\bigg) \mu_{\mA,B}(d\ma,\,db), $$
{\color{black}
we start by applying the Markov inequality with $x\to x^\alpha$ resp. $x \to x^\beta$ with $\beta$ as above to both probabilities:
\begin{align}
\label{eq:b1} {\mathbb P} \bigg(
|{\blue\matrix{\daleth}}_j x| \ge \frac{ e^{n_t q + C_0 +m}}{ |b| e^{ (n-j)\delta}} \bigg) ~&\le~ \frac{\E \big[\abs{{\blue\matrix{\daleth}}_j x}^\alpha \big] \abs{b}^\alpha e^{\alpha(n-j)\delta} }{e^{\alpha(n_t q + C_0 +m)}} ~\le~ \frac{k(\alpha)^j \abs{b}^\alpha e^{\alpha(n-j)\delta} }{c_\alpha e^{\alpha(n_t q + C_0 +m)}} \\
{\mathbb P} \bigg( \|\mPi_{n-j-1}\|> \frac{|b|e^{ (n-j)\delta}}{\|\ma\| e^{C_0+ m + 1}}
\bigg) ~&\le~ \frac{\E \big[ \norm{\mPi_{n-j-1}}^\beta \big] \norm{\ma}^\beta \, e^{\beta(C_0 + m +1)}}{\abs{b}^\beta e^{\beta(n-j)\delta}} ~\le~ \frac{k(\beta)^{n-j} \norm{\ma}^\beta \, e^{\beta(C_0 + m +1)}}{c_\beta k(\beta) \, \abs{b}^\beta e^{\beta(n-j)\delta}} ,
\end{align}
were we used as before Corollary \ref{cor:ksEs} to obtain the second inequalities. Hence, with $\xi=\alpha - \beta$ as before, \begin{align}
II ~&\le~ \sum_{j<n} \sum_{m\ge 0} \frac{k(\alpha)^n e^\beta}{c_\alpha c_\beta k(\beta) e^{\xi{C_0}} e^{\alpha n_t q}} \, e^{-\xi m} \, \left( \frac{e^{\xi \delta} k(\beta)}{k(\alpha)}\right)^{n-j} \, \int \1[\Theta^c](b) \, \norm{\ma}^\beta \abs{b}^\xi \, \mu_{\mA,B}(d\ma,\,db) \\ \label{eq:bb3}
&\le~ \frac{D k(\alpha)^n}{e^{\xi C_0} e^{\alpha n_t q}} \left( \sum_{m \ge 0} e^{-\xi m} \, \sum_{j <n} \left( \frac{e^{\xi \delta} k(\beta)}{k(\alpha)}\right)^{n-j} \, \int \1[\Theta^c](b) \, \norm{\ma}^\beta \abs{b}^\xi \, \mu_{\mA,B}(d\ma,\,db) \right)
\end{align}
with $D= e^\beta /(c_\alpha c_\beta k(\beta))$.
Recall from \eqref{eq:defTheta} that the set $\Theta^c$ is defined in terms of $m$ and $j$, thus in order to resolve the sums, we first have to deal with the integral.
We will apply the H\"older inequality twice. Choose (a small) $\gamma >1$ such that \begin{equation} \label{eq:gam}\E \norm{\mA}^{\gamma \alpha} < \infty \quad \text{ and } \quad\E \abs{B}^{\gamma \alpha} < \infty \quad \text{ and (still) } \quad e^{\xi \delta (1 + \frac{\gamma-1}{\gamma})} k(\beta) < k(\alpha).\end{equation} This is possible due to \eqref{eq:exi} and the moment assumptions \eqref{eq:moment conditions2}. Then, using H\"older first with $p_1=\gamma$, $p_2=\gamma/(\gamma-1)$ and subsequently with $p_1'=\alpha/\beta$, $p_2'=\alpha/\xi$, we obtain
\begin{align}
\label{eq:bb2}\E \1[\Theta^c](B) \norm{\mA}^\beta \abs{B}^\xi ~&\le~ \left(\P{B \in \Theta^c} \right)^{\frac{\gamma-1}{\gamma}} \, \left( \E \norm{\mA}^{\gamma \beta} \abs{B}^{\gamma \xi} \right)^{\frac{1}{\gamma}} \\ \nonumber
~&\le~ \left(\P{B \in \Theta^c} \right)^{\frac{\gamma-1}{\gamma}} \, \left( \E \norm{\mA}^{\gamma \alpha} \right)^{\frac{\beta}{\alpha \gamma}} \, \left( \E \abs{B}^{\gamma \alpha} \right)^{\frac{\xi}{\alpha \gamma}}
\end{align}
Now we apply the Markov inequality with $x \mapsto x^\xi$ to estimate $(\P{B \in \Theta^c})$
\begin{align}
\label{eq:pbtc} \P{\abs{B} > e^{(n_t-j)q + C_0 - (n-j)\delta + m} } ~&\le~ \big( \E \abs{B}^\xi \big) e^{ - {\xi q}(n_t -j) - {\xi}(C_0 + m) } e^{ {\xi \delta} (n-j) } \\
\nonumber ~&\le~ \big( \E \abs{B}^\xi \big) e^{- \xi q \sqrt{n_t}} \, \cdot 1 \cdot e^{\xi \delta (n-j)},
\end{align}
where the last inequality is valid for $j \le n$ (only such $j$ appear in the sum) and follows from the condition $ n < n_t - \sqrt{n_t}$.
Using \eqref{eq:pbtc} in \eqref{eq:bb2} and this in \eqref{eq:bb3}, we obtain
\begin{align}
II ~&\le~ \frac{D k(\alpha)^n}{e^{\xi C_0} e^{\alpha n_t q}} \left( \sum_{m \ge 0} e^{-\xi m} \, \sum_{j <n} \left( \frac{e^{\xi \delta} k(\beta)}{k(\alpha)}\right)^{n-j} \frac{\big( \E \abs{B}^\epsilon \big)^{\frac{\gamma-1}{\gamma}} \left( \E \norm{\mA}^{\gamma \alpha} \right)^{\frac{\beta}{\alpha \gamma}} \, \left( \E \abs{B}^{\gamma \alpha} \right)^{\frac{\xi}{\alpha \gamma}}}{e^{\xi q\frac{ \gamma -1}{\gamma} \sqrt{n_t}}} e^{\xi \delta\frac{ (\gamma -1)}{\gamma} (n-j)} \right) \\
&=~ \frac{D' k(\alpha)^n}{e^{\xi C_0} e^{\alpha n_t q}} \frac{1}{e^{\xi q \frac{\gamma -1}{\gamma} \sqrt{n_t}}} \left( \sum_{m \ge 0} e^{-\xi m} \, \sum_{j <n} \left( \frac{e^{\xi \delta (1+ \frac{\gamma -1}{\gamma})} k(\beta)}{k(\alpha)}\right)^{n-j}
\right)
\end{align}
for a finite constant $D'$, independent of $n, t, C_0$. Recalling \eqref{eq:gam}, both sums converge. Finally, up to a constant, the second quotient can be replaced by $1/\sqrt{n_t}$, and thus we arrive at
\begin{equation}\label{eq:estimateII}
II ~\le~ \frac{D''}{e^{\xi C_o}} \frac{k(\alpha)^n}{\sqrt{n_t} e^{\alpha n_t q}}
\end{equation}
\medskip
\Step[5]: Recall that our original aim was to prove \eqref{aim} with an arbitrarily small $\vartheta$. From the previous steps, we have the estimate
$${\mathbb P}_x \bigg( \bigcup_{j<n} (U_{n,t}\cap W_{j,{n,t}}) \bigg) ~\le~ I + II ~\le~ \frac{C''' + D''}{e^{\xi C_0}} \frac{k(\alpha)^n}{\sqrt{n_t} \, e^{\alpha n_t q}},$$
where $C'''$ and $D''$ are finite constants, independent of $C_0$, which is still a free parameter. Thus, by choosing appropriately large $C_0$, we obtain the assertion.
}
\end{proof}
{\color{black} \begin{lem}\label{lem:sumineq}
Let $0 < \rho <1$, then there is $D_3 < \infty$ such that for all $n \in \N$,
$$ \sum_{j=0}^n \frac{1}{\sqrt{j}} \, \rho^{n-j} ~\le~ D_3 \frac{1}{\sqrt{n}}$$
\end{lem}
\begin{proof}
As the first step, we relabel the sum to $\sum_{j=0}^n \frac{1}{\sqrt{n-j}} \rho^{j}$. Then we split the sum at $J_n:= 1/2 (\log n)/\log (\rho)$ and use that $\rho^{J_n}=1/\sqrt{n}$ and that $n /(n-J_n)$ converges to 1 as $n$ goes to infinity:
\begin{align*}
\sum_{j=0}^{J_n} \frac{1}{\sqrt{n-j}} \, \rho^{j} + \sum_{j={J_n}}^n \frac{1}{\sqrt{n-j}} \, \rho^{j} ~&\le~ \frac{1}{\sqrt{n}} \left( \sum_{j=0}^{J_n} \sqrt{\frac{n}{n- J_n}} \, \rho^j \right) + \rho^{J_n} \left( \sum_{j={J_n}}^n \frac{1}{\sqrt{j}} \, \rho^{j-J_n} \right) \\
&\le~ \frac{1}{\sqrt{n}} \left( \left[ \sup_{n \in \N} \sqrt{\frac{n}{n- J_n}} \right] \sum_{j=0}^{\infty} \rho^j \right) + \frac{1}{\sqrt{n}} \left( \sum_{j={0}}^\infty \rho^{j} \right) ~=: \frac{D_3}{\sqrt{n}}
\end{align*}
\end{proof}
}
\subsection{Finishing the proof}
\begin{proof}[Proof of Lemma \ref{lem:finallemma}]
{\color{black} \Step[1]: We have to prove that
$
{\mathbb P}_x ( \bigcup_{n \in \N} \widetilde V_{n,t} ) \ge \eta t^{-\alpha}
$ for some $\eta >0$ and all large $t$. In order to do so, we can estimate the probability from below by $\P[x]{\bigcup_{n \in K_t} \widetilde V_{n,t}}$, where $K_t$ can be any subset of $\N$, and may depend on $t$.
Applying the inclusion-exclusion formula and Lemma \ref{lem:pos:12.3}, we obtain
\begin{align}
\nonumber {\mathbb P}_x \bigg( \bigcup_{n\in K_t}\widetilde V_{n,t} \bigg) ~\ge &~ \sum_{n\in K_t} {\mathbb P}_x \big({\widetilde V_{n,t}} \big) - \sum_{n,n'\in K_t:\; n>n'}{\mathbb P}_x \big( \widetilde V_{n,t} \cap \widetilde V_{n',t} \big)\\ \label{pvd}
~\ge &~ \kappa_0 \sum_{n\in K_t} {\mathbb P}_x \big({ V_{n,t}} \big) - \sum_{n,n'\in K_t:\; n>n'}{\mathbb P}_x\big( V_{n,t} \cap V_{n',t} \big) ,
\end{align}
where we also used that $(\widetilde V_{n,t} \cap \widetilde V_{n',t}) \subset (V_{n,t} \cap V_{n',t})$, cf. their definitions in \eqref{v1} -- \eqref{v3}.
In order to make the second sum small, we will consider the following specific subsets $K_t$,
\begin{equation}
\label{eq:pos:4}
K_t= \big\{ kC_1:\; n_t-\sqrt{n_t} < kC_1 < n_t - \sqrt{n_t}/2 \big\},
\end{equation}
where the parameter $C_1$ will be chosen later.
\medskip
\Step[2]: In this step, we compute $\P{V_n,t \cap V_{n',t}}$ for $n,n' \in K_t$ with $n>n'$.
Let ${\blue\matrix{\daleth}}_{n'+1}^n = \mM_n^\top \dots \mM_{n'+1}^\top$.
\begin{align}
\P[x]{V_{n,t}\cap V'_{n,t}} ~\le&~ \P{ |{\blue\matrix{\daleth}}_{n'} x|\ge t \mbox{ and } |{\blue\matrix{\daleth}}_n x|\ge t} \nonumber\\
~\le&~ \sum_{m=0}^{\infty} \P{te^m \le |{\blue\matrix{\daleth}}_{n'} x| < t e^{m+1} \mbox{ and } \|{\blue\matrix{\daleth}}_{n'+1}^n\||{\blue\matrix{\daleth}}_{n'} x| >t }\nonumber \\
~\le&~ \sum_{m=0}^{\infty} \P{|{\blue\matrix{\daleth}}_{n'} x| \ge t e^{m}} \P {\|\mPi_{n-n'}^\top\| > e^{-m-1} } \label{vv1}
\end{align}
For the first probability, we can apply the estimate \eqref{eq:br1}: Rewriting it as $\QP[x]{S_{n'} > \log t + m}$,
we have since $n' \in K_t$ that $n' \le n_t= \lceil \log t/q \rceil $ and thus $u:=\log t +m \ge n' q$. Hence,
$$ \P{|{\blue\matrix{\daleth}}_{n'} x| \ge t e^{m}} ~\le~ \frac{C k(\alpha)^{n'}}{\sqrt{n'} \, t^\alpha e^{\alpha m}} ~\le ~ \frac{C' \cdot }{\sqrt{n_t} \, t^\alpha e^{\alpha m}}.$$
For the second inequality, we used that $n' \in K_t$ is comparable to $n_t$.
For the second probability, we use the Markov inequality with $x \to x^\beta$, where $\beta < \alpha$ is such that $k(\beta) < k(\alpha)=1$ (as above), this is possible since $k'(\alpha)>0$. Together with Corollary \ref{cor:ksEs}, we obtain
$$ \P {\|{\blue\matrix{\daleth}}_{n-n'}\| > e^{-m-1} } ~\le~ e^{\beta(m+1)}{\E \norm{\mPi_{n-n'}}^\beta} ~\le~ e^\beta c_\beta^{-1} k(\beta)^{n-n'} \, e^{\beta m}.$$
Using these estimates in \eqref{vv1}, we infer
\begin{align*}
\P[x]{V_{n,t} \cap V_{n',t}} ~\le~ \frac{C' e^\beta}{c_\beta} \frac{ k(\beta)^{n-n'}}{\sqrt{n_t} t^\alpha } \sum_{m=0}^\infty e^{-(\alpha - \beta)m} ~\le~ \frac{C'' k(\beta)^{n-n'}}{\sqrt{n_t} \, t^\alpha}
\end{align*}
\medskip
\Step[3]: Using the previous step and the estimate from Lemma \ref{lem:estimate Vn} (again $k(\alpha)=1$) in \eqref{pvd}, we obtain
\begin{align*}
{\mathbb P}_x \bigg( \bigcup_{n\in K_t}\widetilde V_{n,t} \bigg) ~\ge&~
\kappa_0\sum_{n\in K_t} \frac{D_1}{\sqrt{n_t} e^{\alpha n_t q}} -
\sum_{n\in K_t} \sum_{n'\in K_t:\; n'<n} \frac{C''}{\sqrt{n_t}t^{\alpha}} \cdot k(\beta)^{n-n'}\\
\ge&~ \frac{|K_t|}{\sqrt{n_t} t^{\alpha}}\big( \kappa_0 D_1 - C'' \sum_{n' \in K_t :\; n' < n} k(\beta)^{n-n'} \big)
\end{align*}
Now we use that the cardinality $\abs{K_t} \ge \frac{\sqrt{n_t}}{2 C_1} -1 \ge \frac{\sqrt{n_t}}{4 C_1}$ for all sufficiently large $t$, and that $n - n' \ge C_1$ for $n, n' \in K_t$, $n >n'$.
\begin{align*}
{\mathbb P}_x \bigg( \bigcup_{n\in K_t}\widetilde V_{n,t} \bigg)
\ge&~ \frac{1}{4 C_1}\left( \kappa_0 D_1 - C'' k(\beta)^{C_1} \, \sum_{j=0}^\infty k(\beta)^{j} \right) \frac{1}{t^\alpha} ~=~ \frac{1}{4 C_1}\left( \kappa_0 D_1 - \frac{C'' k(\beta)^{C_1}}{1-k(\beta)} \right) \frac{1}{t^\alpha}
\end{align*}
Since $k(\beta)<k(\alpha)=1$, we can now choose $C_1$ such that the term in the brackets becomes positive. Then, for all $t$ sufficiently large (in particular, such that $K_t$ is nonempty),
$${\mathbb P}_x \bigg( \bigcup_{n\in K_t}\widetilde V_{n,t} \bigg) ~\ge~ \eta t^{-\alpha},$$
with $\eta>0$ being independent of $t$.
}
\end{proof}
|
\section{Introduction}
While a discrepancy between results on determination of the electric
charge radius of the proton has lately attracted attention of theoreticians
and experimentalists, a controversy in
determination of the magnetic radius is rather in shadow. The
situation is summarized in Fig.~\ref{fig:re}. The
proton charge radius has already been discussed in \cite{efit}, which
is referred here as paper I. This paper is a direct continuation of paper I
and we do not reproduce here any plots
or equations from there.
\begin{figure}[thbp]
\begin{center}
\resizebox{0.90\columnwidth}{!}{\includegraphics{rerm.eps} }
\end{center}
\caption{Determination of the rms proton charge and magnetic radii.
(Sick \cite{sick} has evaluated all the world data, but MAMI results
\cite{mami}. Other evaluations of those data produced similar
results (see, e.g., \cite{zhan}).) Ellipses for
electron-proton scattering should be somewhat turned from a pure
horizontal position because of a small correlation between $R_E$ and
$R_M$. For details see \cite{my_adp,my_ufn}.}
\label{fig:re}
\end{figure}
A stronger interest to the situation with the electric charge radius $R_E$
is due to a broader variety of the data and more important
applications, such as determination of the Rydberg constant. While
in the case of the magnetic radius $R_M$ there is a discrepancy
between two scattering results \cite{mami,sick}, the set of results
for $R_E$ also includes the spectroscopic data on hydrogen and
deuterium \cite{codata2010} and on muonic hydrogen
\cite{Nature,Science}.
The contradiction between different values of
$R_M$ is rather serious, while reading the published results literally.
To certain extent it is expected that the discrepancy is partly due to
different treatment of the proton polarizability
contribution~\cite{comm,resp}. However, that may remove only a part of
the discrepancy.
The spectroscopic data and, in particular, results on the
hyperfine-structure (HFS) interval in hydrogen ($1s$) and muonic hydrogen
($2s$) \cite{Science}, may present a source for an independent extraction
of $R_M$, however, at present there is no model-independent constraints
on $R_M$ from the HFS interval in muonic or ordinary hydrogen.
Values published from time to time are deduced from models of the proton
form factors, but there has been no realistic model of the proton
developed to date.
Any comparison of `pure' QED theory with the experimental results on
hydrogen has been `contaminated' for decades by the presence of certain proton
finite-size and polarizability contributions. While the experimental value
of the $1s$ interval in hydrogen had been for a while among the most
accurately measured physical quantities, the most sensitive QED
tests for the HFS-interval theory has been performed not
with the $1s$ interval in hydrogen, but with quantities free of the
influence of the nuclear structure. Such quantities have been provided
by a study of leptonic atoms, such as muonium or positronium. Another
opportunity is a comparison of the $1s$ and $2s$ HFS intervals
measured with the same atom. Details of those QED tests with the HFS
can be found, e.g., in review \cite{my_rep}.
Here we explore a related question. The main purpose of this note is
to estimate the constraints on the magnetic radius of the proton, $R_M$,
from the hyperfine splitting in muonic and ordinary hydrogen. If the proton
polarizability contributions are known with a sufficient accuracy, we
can experimentally determine the value of the proton finite-size
contribution by a comparison of the theory and experiment. Such a
contribution must be sensitive to the distribution of both electric charge
and magnetic moment inside the proton.
Considering that contribution in an appropriate way, we intend to extract a
constraint on a certain combination of $R_E$ and $R_M$.
While the QED effects are well understood (see, e.g.,
\cite{my_rep}), the total theoretical accuracy for the HFS interval in
both muonic and ordinary hydrogen is
completely determined by the proton-structure terms, namely, by the
elastic two-photon contribution and by the proton polarizability
correction. In case of hydrogen the experimental uncertainty is
negligible, while for $\mu$H it is compatible with and somewhat
higher than the theoretical one.
As for calculation of the elastic term, its dominant part can be found
in the external field approximation. We have to deal with integral
\begin{eqnarray}\label{i1def}
I_1^{\rm EM}&\equiv& \int_0^\infty
{\frac{dq}{q^2}}\left[\frac{G_E(q^2)G_M(q^2)}{\mu_p}-1\right]\;,\label{def:i1}
\end{eqnarray}
which determines the dominant proton-finite-size contributions into the HFS
interval in ordinary and muonic hydrogen
\begin{eqnarray}
\Delta E_{\rm HFS}(ns) &=&\frac{8(Z\alpha)m_r}{\pi n^3} E_F
\,I_1^{\rm EM}
\;,
\end{eqnarray}
where $E_F$ is the so-called Fermi energy, $m_r$ is the reduced mass
of a bound electron (in hydrogen) or muon (in muonic hydrogen) and
$\mu_p=2.7928...$ is the proton magnetic moment in units of the
nuclear magnetons. For available experimental data $n=1$ for
ordinary hydrogen\footnote{The $2s$ HFS interval in hydrogen is also
well measured \cite{2s}. The experimental accuracy is worse than for the
$1s$, however, it still supersedes the theoretical accuracy. The $1s$
and $2s$ data are consistent and a separate consideration of the
$2s$ HFS interval would not add any new information on the proton
structure. Meanwhile, comparison of the $1s$ and $2s$ results allows a
sensitive test of QED (see, e.g., \cite{my_rep}).} (see, e.g., a
summary on the $1s$ HFS interval in \cite{1s}) and $n=2$ for muonic
hydrogen \cite{Science}. The other notation used for the integral
under question presents it in terms of the so-called {\em Zemach radius\/}
(or {\em the first Zemach momentum\/})
\begin{eqnarray}
\langle r \rangle_Z &=&-\frac{4}{\pi}I_1^{\rm EM}\;.
\end{eqnarray}
We have no direct experimental knowledge on the integrand in
(\ref{i1def}), which consists of the subtracted form factors of the proton,
${G_E(q^2)G_M(q^2)}/{\mu_p}-1$. In particular, the accurate data
fail at low momenta, which essentially contribute to the integral.
Everything used in the integrand was a result of certain fitting rather than
direct measurements. (We in part explore here ideas
presented previously in \cite{pla} and developed in paper I.)
The situation with the integrand in (\ref{i1def}) is illustrated in
Fig.~\ref{fig:ind1}, where various fractional contributions to the
integrand are estimated from the dipole model and presented as a
function of $q/\Lambda$. The red dot-dashed line is for the subtraction term
with unity. The blue solid line is for the $G_EG_M$ term, which is related to
the data. The integral is fast convergent at high $q$. At low
$q$, say below $0.3\Lambda$, the data contribution produces a large
uncertainty and any successful result for the Zemach contribution
obtained previously was based on a certain, sometimes unrealistic, model.
\begin{figure}[thbp]
\begin{center}
\resizebox{0.8\columnwidth}{!}{\includegraphics{i1integrand1.eps}}
\end{center}
\caption{Fractional contributions to the integrand in (\ref{def:i1})
as a function of $q/\Lambda$ as follows from the dipole model. The red
dot-dashed line is the subtraction term with unity and the blue solid line
is the $G_EG_M$ term, i.e. for the data (cf. \cite{pla,efit}).}
\label{fig:ind1}
\end{figure}
We are going to split the integration into two parts:
\begin{equation}
I=\int_0^\infty {{dq}{...}}\equiv I_<+I_>\equiv
\int_0^{q_0}{{dq}{...}} + \int_{q_0}^\infty {{dq}{...}}
\end{equation}
which are to be treated differently (cf. \cite{pla,efit}).
For higher momenta, we will use direct experimental data (or rather their
realistic approximation). The
accuracy of the form factors is roughly 1\%. The integral over the
direct data is indeed singular at $q_0\to0$, because the
experimental values of $G_E(0)$ and $G_M(0)/\mu_p$ are not equal to
unity exactly --- they are only consistent with unity within the
uncertainty, which produces the singularity. The smaller is $q_0$
the larger is the uncertainty of the related integral.
On the other hand, we can expand the form factors at low momentum
\begin{equation}\label{g2}
\frac{G_E(q^2)G_M(q^2)}{\mu_p}= A + B q^2 + C q^4 + ...\nonumber\\
\end{equation}
Some contributions into $I_{1<}^{\rm EM}$ vanish because of the subtraction and
the uncertainty comes from the remaining terms. The smaller is $q_0$
the smaller is the uncertainty. Here, $A=1$ and
$B=-(R_E^2+R_M^2)/6$. The $B$ contribution to $I_{1<}^{\rm EM}$
\begin{equation}
I_1^{\rm
R}=-\frac{R_E^2+R_M^2}{6}q_0
\end{equation}
is to be treated separately. That is the `signal' that we use to
constrain $R_E^2+R_M^2$. The
leading remaining term is the $C$ term, which is responsible for
the uncertainty.
The idea is to apply a certain model to estimate the uncertainties
and to find a value of $q_0$, which corresponds to the smallest
uncertainty possible (cf. paper I).
Concluding on the model to estimate the uncertainty, we note that
the dipole form factor is a reasonable estimation for the form
factors as far as we discuss general features, but not any accurate
particular value. So, we can, e.g., set for $G_EG_M/\mu_p$
\begin{eqnarray}\label{def:b}
C&=&bC^{\rm dip}\;,\nonumber\\
C^{\rm dip}&=&\frac{10}{\Lambda^4}\;,
\end{eqnarray}
and estimate the $b$ coefficient as $b=1\pm1$ (cf. \cite{efit}).
Here, we use for various preliminary estimations the standard dipole model
\[
G_{\rm dip}(q^2)=\left(\frac{\Lambda^2}{q^2+\Lambda^2}\right)^2
\]
and apply for numerical evaluations $\Lambda^2=0.71\,{\rm GeV}^2$,
which corresponds to $R_{\rm dip}=0.811\;$fm.
\section{Consideration within the dipole model}
Let us perform an evaluation of $I_1^{\rm EM}$ following the consideration of
$I_3^{\rm E}$ in paper I.
The complete dipole value useful for further estimation of the
fractional uncertainties is
\begin{eqnarray}
I_1^{\rm dip}&=& \int_0^\infty {\frac{dq}{q^2}}\left[\left(G_d(q^2)\right)^2-1\right]\nonumber\\
&=&-\frac{35}{32} \frac\pi\Lambda\nonumber\\
&\simeq&-4.08\;{\rm GeV}^{-1}\nonumber\\
&\simeq& -0.805\;{\rm fm}\;.\label{eq:dip}
\end{eqnarray}
\section{Splitting the integral into parts}
As we intend to split the integral into two parts, let us start with
the higher-momentum part
\begin{eqnarray}
I_{1>}^{\rm EM}&=& \int_{q_0}^\infty {\frac{dq}{q^2}}\left[\frac{G_E(q^2)G_M(q^2)}{\mu_p}-1\right]\nonumber\\
&=& \int_{q_0}^\infty {\frac{dq}{q^2}\frac{G_E(q^2)G_M(q^2)}{\mu_p}}
- \frac{1}{q_0}\;.
\end{eqnarray}
Its uncertainty is estimated, by considering the part of the integral,
singular at the limit $q_0\to0$. The result is
\begin{eqnarray}\label{eq:dgg}
\delta I_{1>}^{\rm EM}&=& \delta \int_{q_0}^\infty {\frac{dq}{q^2}}\frac{G_E(q^2)G_M(q^2)}{\mu_p}\nonumber\\
&\simeq& \delta \int_{q_0}^\infty {\frac{dq}{q^2}}\frac{G_E(q_0^2)G_M(q_0^2)}{\mu_p}\nonumber\\
&=& \frac{1}{\nu \Lambda}\frac{2\delta G(q_0^2)}{G(q_0^2)}
\left(\frac{1}{1+\nu^2}\right)^4
\end{eqnarray}
or
\[
\frac{\delta I_{1>}^{\rm EM}}{-I_1^{\rm dip}}\simeq
\frac{0.0058}{\nu}\left(\frac{1}{1+\nu^2}\right)^4\;,
\]
where $\nu=q_0/\Lambda$ and we suggest for our estimations that
both electric and magnetic form factors roughly follow the
standard dipole fit and we experimentally know both of them within
1\% uncertainty
\[
\frac{\delta G(q_0^2)}{G(q_0^2)}\equiv\frac{\delta
G_E(q_0^2)}{G_E(q_0^2)}\simeq\frac{\delta
G_M(q_0^2)}{G_M(q_0^2)}\simeq1\%\;.
\]
Here we apply the dipole values for estimation of absolute and
fractional uncertainties (cf. paper I).
Meanwhile, at low momenta, we find
\begin{eqnarray}
I_{1<}^{\rm EM}&=& \int_0^{q_0} {\frac{dq}{q^2}}\left[\frac{G_E(q^2)G_M(q^2)}{\mu_p}-1\right]\nonumber\\
&=& -\frac{R_E^2+R_M^2}{6}q_0 + \frac{10}{3} b
\frac{q_0^3}{\Lambda^4}\;,
\end{eqnarray}
where $b$ was defined above.
\section{The extraction: a general consideration}
Combining an experimental value, QED contributions and a polarizability
correction we obtain
\begin{eqnarray}
I_1^{\rm exp}&=&\frac{\pi n^3}{8(Z\alpha)m_rE_F} \left( E_{\rm
HFS}^{\rm exp}-E_{\rm HFS}^{\rm QED}-\Delta E_{\rm HFS}^{\rm
polarizability}
\right)\nonumber\\
&=&\frac{\pi}{8(Z\alpha)m_r} \frac{\left( E_{\rm HFS}^{\rm
exp}-E_{\rm HFS}^{\rm QED}-\Delta E_{\rm HFS}^{\rm polarizability}
\right)}{E_{\rm HFS}}
\end{eqnarray}
where we noted that $E_{\rm HFS}\simeq E_F/n^3$ with accuracy sufficient
for the denominator. Alternatively, we can write
\begin{equation}
\langle r \rangle_Z^{\rm exp} =-\frac{1}{2(Z\alpha)m_r} \frac{\left(
E_{\rm HFS}^{\rm exp}-E_{\rm HFS}^{\rm QED}-\Delta E_{\rm HFS}^{\rm
polarizability} \right)}{E_{\rm HFS}}\;.
\end{equation}
Indeed, there are also some higher-order proton-structure
corrections, such as a recoil part of the two-photon exchange. We
assume that they are included if necessary in the QED or
polarizability term.
Often in some papers, they present $\langle r \rangle_Z^{\rm
exp}$ rather than $I_1^{\rm exp}$. Some `experimental' values of $I_1^{\rm
exp}$ are summarized in Table~\ref{t:i1exp}. Any `experimental'
value is a result of an extraction procedure that deeply involves
theory and, in particular, a calculation of the proton polarizability,
which dominates in the uncertainty budget for hydrogen and produces an
uncertainty comparable with the measurement uncertainty for muonic
hydrogen.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|c|c|c|c|c}
\hline
Atom & State & $\langle r \rangle_Z$ & $I_1^{\rm exp}$ & $\delta I_1^{\rm exp}/I_1^{\rm exp}$ & Ref. \\[0.8ex]
\hline
H & $1s$ & 1.047(16) fm & $-4.17(6)\,{\rm GeV}^{-1}$ &1.5\% & \cite{h30} \\[0.8ex]
H & $1s$ & 1.037(16) fm & $-4.13(6)\,{\rm GeV}^{-1}$ & 1.5\% & \cite{h31} \\[0.8ex]
$\mu$H & $2s$ & 1.082(37) fm & $-4.31(15)\,{\rm GeV}^{-1}$ &3.4\% & \cite{Science} \\[0.8ex]
\hline
\end{tabular}
\caption{Some `experimental' values for $I_1$. For numerical evaluations
in this paper we use for hydrogen the value from \cite{h30}.
\label{t:i1exp}}
\end{center}
\end{table}
Meantime, according to our theoretical consideration
\begin{eqnarray}
I_1^{\rm th}&=&-\frac{R_E^2+R_M^2}{6}q_0 + \frac{10}{3}
\bigl(1\pm1\bigr)
\frac{q_0^3}{\Lambda^4}+I_{1>}^{\rm EM}\nonumber\\
&=&-\frac{R_E^2+R_M^2}{6}q_0 +\left(I_1^{\rm EM}-I_1^{\rm R}\right)\;,
\end{eqnarray}
and thus we arrive at
\[
R_E^2+R_M^2=-\frac{6}{q_0}\Bigl(I_1^{\rm exp}-\left(I_1^{\rm EM}-I_1^{\rm
R}\right)\Bigr)\;.
\]
\section{The extraction: the uncertainty budget and its optimization}
It may be useful to introduce the fractional uncertainty of
$R_E^2+R_M^2$
\[
\delta_{2R}=\frac{\delta \bigl( R_E^2+R_M^2\bigr)}{R_E^2+R_M^2}\;.
\]
Since roughly $R_E\approx R_M \approx R_{\rm dip}$, a somewhat
different value
\begin{eqnarray}
\delta'_{2R}&=&\frac{\delta \bigl( R_E^2+R_M^2\bigr)}{2R_{\rm
dip}^2}
\end{eqnarray}
is roughly equal to $\delta_{2R}$, but easier to handle. It is sufficient to minimize the uncertainty of determination of $R_E^2+R_M^2$.
It is equal to the rms sum of partial uncertainties, which are
\begin{eqnarray}
\delta'_{\rm exp}&=&\frac{6}{q_0}\frac{I_1^{\rm dip}}{2R_{\rm
dip}^2}\frac{\delta I_1^{\rm exp} }{I_1^{\rm dip}}\;,\nonumber\\
\delta'_{<}&=&\frac{6}{q_0}\frac{I_1^{\rm dip}}{2R_{\rm dip}^2}
\frac{\delta I_{1<}^{\rm EM} }{I_1^{\rm dip}}\;,
\nonumber\\
\delta'_{>}&=&\frac{6}{q_0}\frac{I_1^{\rm dip}}{2R_{\rm dip}^2}
\frac{\delta I_{1>}^{\rm EM}}{I_1^{\rm dip}}\nonumber\;.
\end{eqnarray}
With
\[
R_{\rm dip}^2=\frac{12}{\Lambda^2}
\]
and
\[
\frac{6}{\Lambda}\frac{-I_1^{\rm dip}}{2R_{\rm dip}^2}=0.8590...\;,
\]
we obtain
\begin{eqnarray}
\delta'_{<}&=&0.83\,\nu^2
\nonumber\;,\\
\delta'_{>}&=&\frac{0.0050}{
\nu^2}\left(\frac{1}{1+\nu^2}\right)^4\nonumber\;.
\end{eqnarray}
To find $\delta_{\rm exp}$, we have to utilize the results from
Table~\ref{t:i1exp}
\begin{eqnarray}
\delta'_{\rm exp,\; H}&=&\frac{0.013}{\nu}\;,\nonumber\\
\delta'_{{\rm exp,\;}\mu{\rm H}}&=&\frac{0.031}{\nu}\nonumber\;,
\end{eqnarray}
where for ordinary hydrogen we use the result from \cite{h30}.
With the uncertainty determined, let us consider behavior of the
uncertainty as a function of $\nu=q_0/\Lambda$. All the partial
uncertainties as well as the total one as a function of $\nu$ are
plotted in Fig.~\ref{fig:unc1H} both for ordinary (top) and muonic
(bottom) hydrogen.
\begin{figure}[thbp]
\begin{center}
\resizebox{0.80\columnwidth}{!}{\includegraphics{i1unc1_h.eps}}
\resizebox{0.80\columnwidth}{!}{\includegraphics{i1unc1_muh.eps}}
\end{center}
\caption{The total fractional uncertainty $\delta'_{2R}$ (a black solid line)
as the rms
sum of sources listed above as a function of $\nu=q_0/\Lambda$. The
partial uncertainties are also presented. The top plot is for
ordinary hydrogen, while the bottom one is for $\mu$H. The red dot-dashed lines
are for the uncertainty of the low-momentum part of the $I_1^{\rm EM}$
integral and the blue solid ones are for high-momentum contribution to the
uncertainty budget; the `experimental' uncertainties, which are
different for muonic and ordinary hydrogen (see
Table~\ref{t:i1exp}), are presented with green dashed lines.}
\label{fig:unc1H}
\end{figure}
The optimal values, which minimize the uncertainty, and the partial
contributions to the total uncertainty for those values are
collected in Table~\ref{t:i1budget}.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|c|c|c|c|c|c|c|c}
\hline Atom & Best $\nu$ & Best $q_0$ & $\delta'_{2R}$ &
$\delta'_{\rm exp}$ & $\delta'_{<}$ & $\delta'_{>}$ & {\em Scat}&
{\em Scat}$^*$
\\[0.8ex]
\hline
H & 0.278 & 0.234 GeV & 9.4\%& 4.8\%& 6.4\%& 4.8\% &{\em 1.3\%}&{\em 3.2\%}\\[0.8ex]
$\mu$H & 0.312 & 0.263 GeV & 13.3\%& 9.9\%& 8.1\%& 3.5\% &{\em 0.9\%}&{\em 2.0\%}\\[0.8ex]
\hline
\end{tabular}
\caption{Parameters for evaluation of the fractional of
uncertainties and fractional scatter of the results (see below). The scatter
of the results from a fit to fit is used to control accuracy. It is not
included into the error budget. {\em Scatter\/} is for the scatter of the
results without fits from \cite{bo1994} and {\em scatter\/}$^*$ is for the
scatter including the fits from \cite{bo1994} (see below for detail).
\label{t:i1budget}}
\end{center}
\end{table}
\section{The fits of the proton form factors}
Now we are to find $I_1^{\rm EM}-I_1^{\rm R}$ by integration over the data.
As in paper I, we use for that a certain set of fits.
As an approximation we utilize the fits for the proton form factors
$G_E$ and $G_M$ from
Arrington and Sick, 2007 \cite{as2007}, Kelly, 2004 \cite{kelly},
Arrington et al., 2007 \cite{am2007}, Alberico et al., 2009
\cite{ab2009}, Venkat et al., 2011 \cite{va2011}, and from
Bosted, 1995 \cite{bo1994}. The details of the fits for the electric
form factor are presented in \cite{efit}, while for the magnetic one
they are summarized in Appendix~\ref{s:fit}.
Two of the fits for $G_M$ are with so-called chain fractions, five
are with Pad\'e approximations with polynomials in $q^2$, and one is
a Pad\'e approximation with polynomials in $q$.
As well as in case of a pure electric integral in \cite{efit}, the
fit (\ref{fit:mbo1994}) of Bosted \cite{bo1994} is perfect for the
tests. It is a Pad\'e approximation with polynomials in $q$, not in
$q^2$. It definitely has a low-momentum behavior strongly different
from others.
The fits are quite close to one another and to the standard dipole
parametrization in the area of interest. They are more or less
consistent to each other and to the dipole one (see
Fig.~\ref{fig:mdip}). Comparison of the fits for the magnetic
form factor to the related fit for the electric form factor is also
presented (see, Fig.~\ref{fig:e-m}).
\begin{figure}[thbp]
\begin{center}
\resizebox{0.85\columnwidth}{!}{\includegraphics{i1mdip.eps}}
\resizebox{0.85\columnwidth}{!}{\includegraphics{i1m-dip.eps} }
\end{center}
\caption{Top: Magnetic form factor $G_M(q^2)/\mu_p$ of the proton:
dipole parametrization and the fits. Bottom: Fractional deviation of
fits from the dipole form factor $(G_M/\mu_p-G_{\rm dip})/G_{\rm
dip}$. Horizontal axis: $q\;$[Gev/$c$]. The blue dashed lines are for the chain
fractions, the green dot-dashed lines are for Pad\'e approximations with
$\tau=q^2/4m_p^2$ and the red solid one is for the Pad\'e approximation
with $q$. The dipole fit in the top graph is presented with a black solid line.}
\label{fig:mdip}
\end{figure}
\begin{figure}[thbp]
\begin{center}
\resizebox{0.85\columnwidth}{!}{\includegraphics{i1e-m.eps}}
\end{center}
\caption{Fractional deviation of the magnetic form factor from the
electric one $(G_E-G_M/\mu_p)/G_E$ (from the same fitting
procedures). Horizontal axis: $q\;$[Gev/$c$]. The blue dashed lines are for the
chain fractions, the green dot-dashed lines are for Pad\'e approximations with
$\tau=q^2/4m_p^2$ and the red solid one is for the Pad\'e approximation
with $q$.}
\label{fig:e-m}
\end{figure}
The low-momentum behavior of the fits is summarized in
Table~\ref{t:exp:mfits}. Indeed, the fit (\ref{fit:mbo1994}) from
\cite{bo1994} is excluded.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|c|c|c|c|c}
\hline
Ref. & ~Fit & Type & $R_E$ & $R_M$ & ~$C$ \\[0.8ex]
&&&[fm]& [fm]&[GeV$^{-4}$]\\
\hline
\cite{as2007}&(\ref{fit:mas2007}) & Chain fraction & 0.90 &0.86& 31.2 \\[0.8ex]
\cite{as2007}& (\ref{fit:mbm2005}) & Chain fraction & 0.90 & 0.87 & 32.2 \\[0.8ex]
\cite{kelly}& (\ref{fit:mkelly}) & Pad\'e approximation ($q^2$) & 0.86 &0.85 & 26.8 \\[0.8ex]
\cite{am2007}&(\ref{fit:mam2007}) & Pad\'e approximation ($q^2$) & 0.88 &0.86& 29.2 \\[0.8ex]
\cite{ab2009}&(\ref{fit:mab20091}) & Pad\'e approximation ($q^2$) & 0.87 &0.87& 28.7 \\[0.8ex]
\cite{ab2009}&(\ref{fit:mab20092}) & Pad\'e approximation ($q^2$) & 0.87 &0.86& 28.2 \\[0.8ex]
\cite{va2011}&(\ref{fit:mva2011}) & Pad\'e approximation ($q^2$) & 0.88 &0.86& 29.5 \\[0.8ex]
\hline
\end{tabular}
\caption{The low-momentum expansion of the fits can be expressed as
$G_E G_M/\mu_p= 1 - (R_E^2+R_M^2) \,q^2/3 + C q^4 +
...$. The values in the Table are given for central values of the
fits without any uncertainty. The references are given to the papers
where both electric and magnetic form factors are presented. The
summary on the applied fits for the electric form factor of the
proton can be found in \cite{efit}. The references to the
equations are given here for the magnetic form factor (see Appendix
\ref{s:fit}). The related values for the standard dipole fit are
$R_E=R_M=0.811\;{\rm fm}$ and $C=19.8\;{\rm
GeV}^{-4}$.\label{t:exp:mfits}}
\end{center}
\end{table}
\section{Integration over the fits}
Integrating over the fits for the optimal $q_0({\rm H})= 0.222\;$GeV
for hydrogen, we find that $I_{1>}^{\rm EM}$ varies from $-2.92\,$GeV$^{-1}$ to
$-2.89\,$GeV$^{-1}$ if we exclude Pad\'e approximation in $q$
\cite{bo1994} or from $-2.97\,$GeV$^{-1}$ if we include it. For
detail see Table~\ref{t:i1:fit}.
Here,
we accept
\[
I_{1>}^{\rm EM}({\rm H})=-2.90(2)\;{\rm GeV}^{-1}
\]
as the mean value (excluding (\ref{fit:mbo1994})), that leads to
\[
I_1^{\rm EM}({\rm H})-I_1^R({\rm H})=-2.82(11)\;{\rm GeV}^{-1}\;.
\]
The uncertainty of integral above does not include scattering in the
calculation of $I_{1>}^{\rm EM}$ because we estimated the uncertainty of
this term in a more conservative way as explained above.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|c|c}
\hline
Fit & Type & $I_{1>}^{\rm EM}$ \\[0.8ex]
\hline
\cite{as2007}& Chain fraction & $-2.91\;{\rm GeV}^{-1}$\\[0.8ex]
\cite{as2007}& Chain fraction & $-2.92\;{\rm GeV}^{-1}$\\[0.8ex
\cite{kelly} & Pad\'e approximation ($q^2$) & $-2.89\;{\rm GeV}^{-1}$\\[0.8ex]
\cite{am2007}& Pad\'e approximation ($q^2$) & $-2.90\;{\rm GeV}^{-1}$\\[0.8ex]
\cite{ab2009}& Pad\'e approximation ($q^2$) & $-2.90\;{\rm GeV}^{-1}$\\[0.8ex]
\cite{ab2009}& Pad\'e approximation ($q^2$) & $-2.90\;{\rm GeV}^{-1}$\\[0.8ex]
\cite{va2011}& Pad\'e approximation ($q^2$) & $-2.90\;{\rm GeV}^{-1}$\\[0.8ex]
\cite{bo1994}& Pad\'e approximation ($q$) & $-2.97\;{\rm GeV}^{-1}$\\[0.8ex]
\hline
\end{tabular}
\caption{Scatter of data for $I_{1>}^{\rm EM}$ for hydrogen at
`optimal' $q_0\simeq 0.278\,\Lambda=0.234\,{\rm
GeV}/c$.\label{t:i1:fit}}
\end{center}
\end{table}
Eventually, we obtain a constraint on the magnetic radius from the HFS
interval in hydrogen
\begin{equation}
\frac{R_E^2+R_M^2}{2R_{\rm dip}^2} = 1.025(94) \;,
\end{equation}
and we remind that for the standard dipole parametrization $R_{\rm
dip}^2=0.658\;{\rm fm}^2$.
The related fractional scatter is 0.013 if we exclude (\ref{fit:mbo1994}) from
the consideration and it is 0.032 if we include it. The result
for the combination of the proton electric and magnetic radius is
consistent with the value from the standard dipole model within its
9\% uncertainty.
The same evaluation can be performed for various $q_0$ and the
results are summarized in Table~\ref{t:un1h:q}. All the results are
consistent. The scatter is below the uncertainty except for very low
$q_0$, where behavior of the fits becomes model dependent.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|c|c|c|c}
\hline
$q_0/\Lambda$ &$q_0$ & $({R_E^2+R_M^2})/{2R_{\rm dip}^2}$ &{\em Scatter} &
{\em Scatter}$^*$ \\[0.8ex]
\hline
0.20 & 0.169\;GeV & 0.99(13) & {\em 0.03} & {\em 0.10} \\[0.8ex]
0.25 & 0.211\;GeV & 1.02(9) & {\em 0.02}& {\em 0.05} \\[0.8ex]
0.30 & 0.253\;GeV & 1.03(10) & {\em 0.01}& {\em 0.02} \\[0.8ex]
0.35 & 0.295\;GeV & 1.04(11) & {\em 0.007}& {\em 0.01} \\[0.8ex]
0.40 & 0.337\;GeV & 1.05(14) & {\em 0.004}& {\em 0.007} \\[0.8ex]
0.50 & 0.421\;GeV & 1.08(21) & {\em 0.002}& {\em 0.002} \\[0.8ex]
\hline
\end{tabular}
\caption{The results for $({R_E^2+R_M^2})/{2R_{\rm dip}^2}$ at
various $q_0$ for hydrogen. \label{t:un1h:q}}
\end{center}
\end{table}
Similar treatment for muonic hydrogen produces $q_0(\mu{\rm H})=
0.263\;$GeV as the optimized value. The results of integration over
the fits for $I_{1>}^{\rm EM}$ vary from $-2.77\:$GeV$^{-1}$ excluding
(\ref{fit:mbo1994}) and from $-2.80\:$GeV$^{-1}$ including it to
$-2.74\:$GeV$^{-1}$ (see Table~\ref{t:i1:fit:mu}). We consider
\[
I_{1>}^{\rm EM}(\mu{\rm H})=-2.75(1)\;{\rm GeV}^{-1}
\]
as the mean value that leads to
\[
I_1^{\rm EM}(\mu{\rm H})-I_1^R(\mu{\rm H})=-2.63(13)\;{\rm
GeV}^{-1}\;.
\]
The uncertainty of integral above does not include scattering in a
calculation of $I_>$ because we estimated the uncertainty of this term
in a more conservative way as explained above.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|c|c}
\hline
Fit & Type & $I_{1>}^{\rm EM}$ \\[0.8ex]
\hline
\cite{as2007}& Chain fraction & $-2.76\;{\rm GeV}^{-3}$\\[0.8ex]
\cite{as2007}& Chain fraction & $-2.77\;{\rm GeV}^{-3}$\\[0.8ex
\cite{kelly}& Pad\'e approximation ($q^2$) & $-2.74\;{\rm GeV}^{-3}$\\[0.8ex]
\cite{am2007}& Pad\'e approximation ($q^2$) & $-2.75\;{\rm GeV}^{-3}$\\[0.8ex]
\cite{ab2009}& Pad\'e approximation ($q^2$) & $-2.76\;{\rm GeV}^{-3}$\\[0.8ex]
\cite{ab2009}& Pad\'e approximation ($q^2$) & $-2.75\;{\rm GeV}^{-3}$\\[0.8ex]
\cite{va2011}& Pad\'e approximation ($q^2$) & $-2.75\;{\rm GeV}^{-3}$\\[0.8ex]
\cite{bo1994}& Pad\'e approximation ($q$) & $-2.80\;{\rm GeV}^{-3}$\\[0.8ex]
\hline
\end{tabular}
\caption{Scatter of data for $I_{1>}^{\rm EM}$ for muonic hydrogen
at `optimal' $q_0\simeq 0.312\Lambda=0.263\,{\rm
GeV}/c$.\label{t:i1:fit:mu}}
\end{center}
\end{table}
The constraint from the HFS interval in muonic hydrogen is found to
be
\begin{equation}
\frac{R_E^2+R_M^2}{2R_{\rm dip}^2}= 1.13(13) \;.
\end{equation}
The fractional scatter is 0.009 (excluding (\ref{fit:mbo1994})) or 0.020
(including (\ref{fit:mbo1994})). The result is consistent with the
value from the standard dipole moment within the uncertainty of
13\%.
The results obtained at various values of the separation parameter
$q_0$ are consistent to each other (see Table~\ref{t:un1muh:q} for details).
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|c|c|c|c}
\hline
$q_0/\Lambda$ &$q_0$ & $({R_E^2+R_M^2})/{2R_{\rm dip}^2}$ &{\em scatter} &{\em scatter}$^*$ \\[0.8ex]
\hline
0.20 & 0.169\;GeV & 1.14(19) & {\em 0.03} & {\em 0.10} \\[0.8ex]
0.25 & 0.211\;GeV & 1.13(14) & {\em 0.02}& {\em 0.05} \\[0.8ex]
0.30 & 0.253\;GeV & 1.13(13) & {\em 0.01}& {\em 0.02} \\[0.8ex]
0.35 & 0.295\;GeV & 1.13(14) & {\em 0.007}& {\em 0.01} \\[0.8ex]
0.40 & 0.337\;GeV & 1.13(16) & {\em 0.004}& {\em 0.007} \\[0.8ex]
0.50 & 0.421\;GeV & 1.14(22) & {\em 0.002}& {\em 0.002} \\[0.8ex]
\hline
\end{tabular}
\caption{The results for $({R_E^2+R_M^2})/{2R_{\rm dip}^2}$ at various
$q_0$ for muonic hydrogen. \label{t:un1muh:q}}
\end{center}
\end{table}
\section{Conclusions}
Our strategy to evaluate $I_1^{\rm EM}$ was dictated by our purpose,
which is to determine $R_M$ (cf. \cite{pla}). For a different purpose
the strategy would be different.
To obtain
a constraint on the magnetic radius of the proton $R_M$, the
compilation of all constraints on the electromagnetic radii of the proton
is presented in
Fig.~\ref{fig:re:new}. We plot there constraints from
Fig.~\ref{fig:re} and also present two constraints on $R_E^2+R_M^2$
derived here from a study of the HFS intervals in ordinary and muonic
hydrogen.
\begin{figure}[thbp]
\begin{center}
\resizebox{0.99\columnwidth}{!}{\includegraphics{r2bnew.eps} }
\end{center}
\caption{Determination of the rms proton charge and magnetic radii.
Notation is the same as in Fig.~\ref{fig:re}. Two new striped belts are
added from the HFS constraints in ordinary and muonic hydrogen
above. The former is filled with vertical lines and the latter is with
the horizontal ones. Their colors are the same as for the Lamb shift
in the related
atomic system. The squared area is the overlap of two striped areas.}
\label{fig:re:new}
\end{figure}
That is a general picture. It already has certain features similar
to those considered in paper I. The overall accuracy of
spectroscopic extractions of the magnetic radius looks comparable
with scattering
results---not with their claimed uncertainty, but with their
discrepancy.
The preliminary results on the proton magnetic radius from atomic
spectroscopy are presented in Table~\ref{t:rm}. It
involves all possible combinations of spectroscopic constraints on
$R_E^2+R_M^2$ (from HFS) and on $R_E$ (from the Lamb shift). Note
that ${\delta R_M}/{R_M}\approx\delta^\prime_{2R}$, assuming that
$R_E$ is known with a good accuracy and that roughly $R_E\approx R_M
\approx R_{\rm dip}$.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|l|l}
\hline
Transition/atom & Lamb ($\mu$H) & Lamb (H) \\[0.8ex]
\hline
HFS (H) & 0.80(8) fm & 0.76(8) fm \\[0.8ex]
HFS ($\mu$H) & 0.88(11) fm & 0.85(11) fm \\[0.8ex]
\hline
\end{tabular}
\caption{Magnetic radius of the proton from combining HFS and the
Lamb shift in muonic and ordinary hydrogen.\label{t:rm}}
\end{center}
\end{table}
If we accept the value of the proton charge radius as
\[
R_E=0.86(2)\;{\rm fm}\;,
\]
which seems a reasonable choice until the controversy in its determination
is not resolved, then we arrive at
\[
R_M=0.78(8)\;{\rm fm}\;,
\]
as to the best constraint on the magnetic radius from spectroscopy.
We have performed a number of consistency checks described above such as
consideration of various values of the separation parameter. All the
results are consistent. The estimation of the $q^4$ term is consistent with
all fits with reasonable behavior at low $q$ discussed in Appendix
(as well as with the fits from MAMI (see \cite{thesis} for detail)).
In case of the fits considered above the only fit with unreasonable behavior
is that from \cite{bo1994}, which in particular produces infinite values
of the charge and mangetic radii. It is important that all fits but
(\ref{fit:mbo1994}) agree with each other within at 1\% level in the region
where the separation parameter was chosen in Tables~\ref{t:un1h:q} and
\ref{t:un1muh:q} as seen in Fig.~\ref{fig:40}.
The fit of \cite{bo1994} with incorrect behavior at low momentum
transfer is responsible for a scatter of values of $G_M(q_0)$ bigger
than $\pm1\%$ (see Fig.~\ref{fig:40}) (cf. \cite{efit}). However,
taking into account its unrealistic behavior, that is acceptable. The
other fits agree with each other within 1\% in the region crucial for a
choice of $q_0$. A similar situation with the electric radius (see
Fig.~6 of paper I).
\begin{figure}[thbp]
\begin{center}
\resizebox{0.85\columnwidth}{!}{\includegraphics{i1m-dip40.eps} }
\resizebox{0.85\columnwidth}{!}{\includegraphics{i1e-m40.eps} }
\end{center}
\caption{Top: Fractional deviation of the magnetic form factor from the
dipole form factor $(G_M/\mu_p-G_{\rm dip})/G_{\rm dip}$. Bottom:
Fractional deviation of the magnetic form factor from the electric
one $(G_E-G_M(q^2)/\mu_p)/G_E$. Horizontal axis: $q\;$[Gev/$c$] and
the range $0.2-0.5\;{\rm GeV}/c$ is crucial for $q_0$ in all three
cases considered.}
\label{fig:40}
\end{figure}
A value of the magnetic radius extracted from electron-proton scattering
strongly depends on
treatment of the proton polarizability \cite{comm,resp}. To apply
the Rosenbluth separation, one has to rely on a certain model for the
proton polarizability. We note, however, that the electric form factor
$G_E$ at low momentum transfer is less
sensitive to the model and as far as we are not going to go too low,
one may use experimental results on $G_M/G_E$ from recoil polarimetry
(see, e.g.,
\cite{zhan}), which are free of the polarizability problem.
Similarly to the case of examination for $R_E$ in \cite{efit}, we
conclude that the estimation of the uncertainty of the form factors
at the level of 1\% for applied $q_0$ is validated by the behavior
of the fits and by the scale of the scatter. Nevertheless, a direct
investigation of the problem would be useful.
The author is grateful to S. Eidelman and V. Ivanov for useful discussions.
This work was supported in part by DFG under grant HA 1457/9-1.
|
\section{Introduction}
The effect of a quenched disorder on critical phenomena is a central topic in equilibrium statistical mechanics. In many cases it is expected that the presence of impurities in a system \textit{rounds} or \textit{smoothes} the phase transition in the following sense: the order parameter can be continuous at the phase transition for the disordered system whereas it presents a discontinuity for the pure system (see e.g. the pioneering work of Imri and Ma \cite{cf:IM}). An instance for which this phenomenon is rigorously proved is the magnetization transition of the two dimensional random field Ising model at low temperature (see \cite{cf:AW}).
\medskip
This phenomenon has been particularly studied for the polymer pinning on a defect line (introduced by Fisher in \cite{cf:Fisher}).
Whereas the model can be defined for a renewal with a distribution tail which is heavier than exponential (see \eqref{rules}), the case of power-law tail has focused most of the attention, due to its physical interpretation and its rich mathematical structure. The interested reader can refer to \cite{cf:GB, cf:GB2, cf:denH} for reviews on the subject. The smoothing of the free-energy curve for the pinning model with power-law tails was proved in \cite{cf:GT05} (with some restriction on the law of the disorder see \cite{cf:CdH} for a recent generalization of the result; see also \cite{cf:BL2, cf:LT} for related models). This confirmed predictions by theoretical physicists \cite{cf:DHV} based on an interpretation of the Harris criterion \cite{cf:Harris}. Some other consequences of the introduction of disorder such as critical point shift were studied in \cite{cf:A06, cf:T08, cf:DGLT07, cf:AZ08, cf:GLT08, cf:GLT09, cf:AZ10}.
\medskip
The present paper aims to study how this phenomenology transposes for renewals with a much lighter tail, stretched exponential ones. Whereas this issue does not seem to be discussed much in the literature, it is clear from a mathematical point of view that the type of argument used in \cite{cf:GT05} do not extend to that case (see Section \ref{smoothpol} for a more detailed discussion). This hints to the fact that when renewal tails gets lighter, Harris predictions on disorder relevance might not apply (or at least not in a straightforward manner). We show that this is indeed the case and provide a necessary and sufficient condition on the return exponent for smoothing of the free-energy curve to hold.
\medskip
Let us notice finally notice that renewals with stretched exponential tails have recently been the object of a study by Torri \cite{cf:T14} with a different perspective: he focuses on the issue of the scaling limit of the process when the environment is heavy tailed.
\subsection{The disordered pinning model}
Let us shortly introduce the model: set $\tau:=(\tau_0,\tau_1,\ldots)$ to be a renewal process of law
${\ensuremath{\mathbf P}} $, with inter-arrival law $K(\cdot)$, {\sl i.e.}, $\tau_0=0$ and
$\{\tau_i-\tau_{i-1}\}_{i\in\mathbb{N}}$ is a sequence of IID positive integer-valued
random variables.
Set
\begin{equation}
K(n):={\ensuremath{\mathbb P}} [\tau_1=n].
\end{equation}
We assume that
\begin{equation}\label{rules}
\lim_{n\to \infty} n^{-1} \log K(n)=0.
\end{equation}
Note that with a slight abuse of notation, $\tau$ can also be considered as a subset of ${\ensuremath{\mathbb N}} $ and we will write
$\{ n\in \tau \}$ for $\{ \exists i, \ \tau_i=n \}$.
The random potential $\omega:=\{\omega_1,\omega_2,\ldots\}$ is a sequence of
IID centered random variables which have unit variance and exponential moments of all order
\begin{equation}\label{defgl}
\lambda(\beta):=\log {\ensuremath{\mathbb E}} [e^{\beta\omega}]<\infty.
\end{equation}
\medskip
Given $\beta>0$ (the inverse temperature) and $h\in {\ensuremath{\mathbb R}} $,
we define ${\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N$ a measure whose Radon-Nikodym derivative w.r.t ${\ensuremath{\mathbf P}} $ is given by
\begin{equation}\label{density}
\frac{\dd {\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N}{\dd {\ensuremath{\mathbf P}} }(\tau):= \frac{1}{Z^{\beta,h,\omega}_N}\exp\left(\sum_{n=0}^N(\beta \omega_n+h)\delta_n\right)\delta_N
\end{equation}
where $\delta_n=\mathbf{1}_{\{n\in \tau\}}$ and $Z^{\beta,h,\omega}_N$ is the renormalizing constant which makes ${\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N$ a probability law:
\begin{equation}
\label{eq:Znh}
Z^{\beta,h,\omega}_{N}:={\ensuremath{\mathbf E}} \left[e^{\sum_{n=1}^N(\beta\omega_n+h)\delta_n}\delta_N\right].
\end{equation}
\begin{remark} \label{rembc} In the definition \eqref{density} of ${\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N$, the $\delta_N$ corresponds to constraining the end point to be pinned.
This conditioning is present for technical reasons and makes some computations easier but is not essential.
\end{remark}
By ergodic super-additivity, (see \cite[Chap. 4]{cf:GB}), the limit
\begin{equation}
\label{eq:F_nh}
\textsc{f}(\beta,h):=\lim_{N\to\infty}\frac1N\log Z^{\beta,h,\omega}_{N}
\end{equation}
exists and is non-random. It is non-negative because of assumption \eqref{rules} and convex in $h$ as a limit of convex functions.
The expectation also converges to the same limit
\begin{equation}\label{freen}
\textsc{f}(\beta,h)=\lim_{N\to\infty}\frac1N{\ensuremath{\mathbb E}} \log Z^{\beta,h,\omega}_{N}.
\end{equation}
The function $\textsc{f}$ is called the free-energy (or sometimes pressure) of the system. Its derivative in $h$ gives the asymptotic contact fraction of the renewal process, i.e.\ the mean number of contact per unit length,
\begin{equation}\label{gro}
\partial_h \textsc{f}(\beta,h)=\lim_{N\to \infty} \frac 1 N {\ensuremath{\mathbf E}} ^{\beta,h,\omega}\left[\sum_{n=1}^N \delta_n\right].
\end{equation}
The above convergence holds by convexity as soon as $\partial_h \textsc{f}(\beta,h)$ is defined (i.e.\ everywhere except eventually at a countable number of points).
If \eqref{rules} holds, the system undergoes a phase transition from a de-pinned state
($\textsc{f}(\beta,h)\equiv 0$) to a pinned one ($\textsc{f}(\beta,h)>0$ and $ \partial_h \textsc{f}(\beta,h)>0$) when $h$ varies.
\medskip
We define $h_c(\beta)$, the critical point at which this transition occurs
\begin{equation}
h_c(\beta):= \min\left\{ h \ | \ \textsc{f}(\beta,h)>0\right\}.
\end{equation}
As the underlying renewal process $\tau$ is recurrent, we have $h_c(0)=0$.
From \cite[Theorem 2.1]{cf:Alea}, the free energy is infinitely differentiable in $h$ on $(h_c(\beta),\infty)$ (so that \eqref{gro} holds everywhere except maybe at the critical point).
The phase transition for the pure system, that is, for $\beta=0$, is very well understood. The pure model is said to be exactly solvable
and there is a closed expression for $\textsc{f}(0,h)$ in terms of the renewal function $K$ (see \cite{cf:Fisher}).
\subsection{Disorder relevance and Harris criterion for power-law renewals}
The disordered system ($\beta>0$) is much more complicated to analyze and has given rise to a rich literature, most of which devoted to the case where when $n\to \infty$
\begin{equation}\label{poly}
K(n)=c_K n^{-(1+\alpha)}(1+o(1))
\end{equation}
for some $\alpha>0$. For the pure model, the free-energy vanishes like a power of $h$ at $0+$ (see \cite[Theorem 2.1]{cf:GB}).
\begin{equation}
\textsc{f}(0,h)= c'_K h^{\max(1,\alpha^{-1})}(1+o(1)),
\end{equation}
for $\alpha\ne 1$ (a logarithmic correction is present in the case $\alpha=1$).
The main question for the study of disordered pinning model is how this property of the phase transition is affected by the introduction of disorder.
For $\beta>0$, does there exist $\nu$ such that at the vicinity of $h_c(\beta)_+$
\begin{equation}\label{nu}
\textsc{f}(\beta,h)\approx (h-h_c(\beta))^{\nu}?
\end{equation}
If this holds, is $\nu$ equal $\max(1,\alpha^{-1})$, like for the pure system?
A first partial answer to that question was given by Giacomin and Toninelli \cite{cf:GT05} (or in \cite{cf:CdH} with more generality) where it was shown that
\begin{equation}
\textsc{f}(\beta,h)\le C\left( \frac{h-h_c(\beta)}{\beta}\right)^{2},
\end{equation}
meaning that the quenched critical exponent for the free-energy $\nu$, if its exists, satisfies $\nu \ge 2$.
In particular it cannot be equal to the one of the pure system when $\alpha>1/2$.
\medskip
One the other hand, for small $\beta$ and $\alpha<1/2$, it was shown by Alexander \cite{cf:A06} (see \cite{cf:T08, cf:Lmart} for alternative proofs) that $h_c(\beta)=-\lambda(\beta)$ (recall \eqref{defgl}) and that when $u\to 0+$
\begin{equation}
\textsc{f}(\beta,u-\lambda(\beta))= \textsc{f}(0,u)(1+o(1))
\end{equation}
meaning that $\nu$ exists and is equal to $\max(1,\alpha^{-1})$ as for the pure model.
\medskip
Another aspect of the relevance of disorder is the shift of the \textsl{quenched} critical point with respect to the \text{annealed} one.
The annealed critical point is the one corresponding to the phase transition of the annealed partition function obtained by averaging over the environment
\begin{equation}\label{annealedbound}
h_c^a(\beta):=\inf \left\{ h \ | \ \lim_{N\to \infty} \frac{1}{N} \log {\ensuremath{\mathbb E}} \left[ Z^{\beta,h,\omega}_{N}\right]>0\right\}=-\lambda(\beta).
\end{equation}
It follows from Jensen's inequality that
\begin{equation}\label{quenchann}
h_c(\beta)\ge h^a_c(\beta).
\end{equation}
The question of whether the above inequality is strict
was investigated in
\cite{cf:DGLT07, cf:AZ08, cf:GLT08, cf:GLT09} yielding the conclusion that $h_c(\beta)>-\lambda(\beta)$ for every $\beta>0$ and $\alpha\ge 1/2$.
\medskip
These results were predicted in the Physics literature \cite{cf:DHV, cf:FLN}, based on an interpretation of the Harris criterion \cite{cf:Harris}: when the specific-heat exponent of the pure system (for the pinning model, this exponent is equal to $2-\max(1,\alpha^{-1})$) is positive, then disorder affects the critical properties of the system and is said to be relevant, whereas when the specific-heat exponent is negative disorder is irrelevant for small values of $\beta$.
\medskip
Relevant disorder affects both the location of the critical point which is shifted with respect to the annealed bound \eqref{quenchann}
\cite{cf:DGLT07, cf:AZ08,cf:GLT08, cf:GLT09}, and the critical exponent of the free-energy \cite{cf:GT05, cf:CdH}.
Note that the value of $\nu$ (and even its existence) when disorder is relevant is an open question even among physicists; let us mention the recent work
\cite{cf:DR} where heuristics in favor $\nu=\infty$ (infinitely derivable free-energy at the critical point) are given for a toy-model.
In this paper, we choose to look at renewal processes whose tails are stretched exponentials, i.e we assume that there exists $\zeta\in(0,1)$ such that
\begin{equation}
K(n)\approx \exp(-n^{\zeta}),
\end{equation}
in some sense.
As the increments of $\tau$ have finite mean, the transition of the pure model is of first order, meaning that $\textsc{f}(0,h)$ is not derivable at $h_c(0)=0$ positive recurrent. More precisely,
from \cite[Th. 2.1]{cf:GB} one has
\begin{equation}
\label{eq:annF}
\textsc{f}(0,h)\stackrel{h\searrow0}\sim \frac{h}{{\ensuremath{\mathbf E}} [\tau_0]}.
\end{equation}
as for the case $\alpha>1$ in \eqref{poly}. Hence a standard interpretation of the Harris criterion would tell us that disorder should be relevant for every $\beta$.
This is partially true in the sense that this conclusion is right if one considers only the question of the critical point shift.
The method developed in \cite{cf:DGLT07} can be adapted almost in a straightforward manner to show that
\begin{proposition}\label{mook}
When $K(n)$ has stretched-exponential tails, then for all $\beta>0$,
\begin{equation}
h_c(\beta)>-\lambda(\beta).
\end{equation}
\end{proposition}
The more challenging question is the one about the order of the phase transition. Indeed the smoothing inequality proved in \cite{cf:GT05} strongly relies of the fact that
$K(\cdot)$ has a power-law tail.
\medskip
We are in fact able to find a necessary and sufficient condition on $\zeta$ for a smoothing inequality to hold: we prove that when $\zeta>1/2$, the transition remains of first order for the disordered system, while for $\zeta\le 1/2$ the transition is rounded.
We also give upper and lower bounds, which do not coincide, on the exponent $\nu$, informally defined in \eqref{nu}, when rounding occurs, in particular we show that for any value of $\zeta\in(0,1)$, the disordered phase transition remains of finite order.
\section{Presentation of the results} \label{sec:nHmodel}
\subsection{Results}
We assume here and in what follows that there exist a constant $c_K$ and $\zeta\in (0,1)$ which is such that
\begin{equation}
\label{eq:K}
K(n)= c_K(1+o(1))\exp(-n^{\zeta}).
\end{equation}
The law $K(n)$ as well as the law of $\omega$ are considered to be fixed, and constants that are mentioned throughout the proof can depend on both.
Unless it is specified, they will not depend on $\beta$ and $h$.
For our first result, we need to assume that the law of our product environment satisfies a concentration inequality.
We say that $F: {\ensuremath{\mathbb R}} ^N\to {\ensuremath{\mathbb R}} $ is Lipschitz if for some $k>0$ if
\begin{equation}
\| F \|_{\mathrm{lip}}=\sup_{x\ne y \in {\ensuremath{\mathbb R}} ^N} \frac{|F(x)-F(y)|}{|x-y|}<\infty
\end{equation}
where $|x-y|= \sqrt{ \sum (x_i-y_i)^2}$ is the Euclidean norm.
\begin{assumption}\label{logsob}
There exist constants $C_1$ and $C_2$ such that for any $N$ and for any Lipschitz convex function $F$ on ${\ensuremath{\mathbb R}} ^N$, one has
\begin{equation}
{\ensuremath{\mathbb P}} \left(
| F(\omega_1,\dots,\omega_N)-{\ensuremath{\mathbb E}} \left[ F(\omega_1,\dots,\omega_N)\right]| \ge u \right) \le C_1e^{-\frac{u^2}{C_2 \| F \|_{\mathrm{lip}}^2}}
\end{equation}
\end{assumption}
A crucial point here is that inequality does not depend of the dimension $N$. This is the reason why we use concentration for the Euclidean norm rather than for the $L_1$ norm.
\begin{remark}
The concentration assumption is not very restrictive, it holds for bounded $\omega$ (see \cite[Chapter 4]{cf:Ledoux}),
or when $\omega$ satisfies a $\log$-Sobolev inequality (see \cite[Chapter 5]{cf:Ledoux} in this case there is no convexity required). This second case includes in particular the case of Gaussian variables and many other classic laws.
\end{remark}
Our first result states that the transition is of first order for the system for $\zeta>1/2$ (no smoothing holds). Here and in what follows $x_+:= \max(x,0)$ denotes the positive part of $x\in {\ensuremath{\mathbb R}} $.
\begin{theorem}\label{mainres}
Assume that Assumption \ref{logsob} holds.
\begin{itemize}
\item[(i)] For $\zeta>1/2$, there exists a constant $c$ such that for all $\beta$ and $h$,
\begin{equation}\label{ground}
\textsc{f}(\beta,h)\ge c\max(1,\beta^{-2}) (h-h_c(\beta))_+.
\end{equation}
\item[(ii)]
For $\zeta\le 1/2,$ there exists a constant $c$ and $u_0(\beta)$ such that for all $u\in(0,u_0)$
\begin{equation}\label{ground2}
\textsc{f}(\beta,h_c(\beta)+u )\ge c\left(\frac{u}{\beta^2 | \log u|}\right)^{\frac{1-\zeta}{\zeta}}.
\end{equation}
\end{itemize}
\end{theorem}
\medskip
Our second result shows that smoothing holds for $\zeta< 1/2$.
For this result we need to assume that the environment is Gaussian.
The assumption could be partially relaxed but the exposition of the Gaussian case is much easier.
Let us mention that the recent work \cite{cf:CdH} gives hopes to extend the proof to general $\omega$.
\begin{theorem}\label{mainres2}
Let us assume that the environment is Gaussian. Then for all $\zeta<1/2$ there exists a constant $c$ (which depends on $K$) such that in a neighborhood of $h_c(\beta)$
\begin{equation}
\textsc{f}(\beta,h)\le c \left(\frac{h-h_c(\beta)}{\beta}\right)^{2(1-\zeta)}_+.
\end{equation}
\end{theorem}
Finally with an extra assumption on $K(\cdot)$ we are able to state that the transition is smooth also when $\zeta=1/2$.
We say that $K(n)$ is $\log$ convex if $\log K$ can be extended to a convex function on ${\ensuremath{\mathbb R}} _+$; or equivalently if one has
\begin{equation}\label{logconvex}
\forall n,l \in {\ensuremath{\mathbb N}} , \ n>l>1 \Rightarrow K(n+1)K(l-1)\ge K(n)K(l).
\end{equation}
This assumption is necessary to prove positive correlation, or the FKG inequality (see \cite{cf:FKG}) for the disordered renewal.
\begin{theorem}\label{mainres3}
Assume that $\log K(n)$ is a convex function of $n$. Then for $\zeta=1/2$ one has
\begin{equation}
\textsc{f}(\beta,h)= o\left((h-h_c(\beta))_+\right).
\end{equation}
\end{theorem}
\begin{remark}
The $\log$-convex assumption is not that restrictive and is rather natural as assumption \eqref{rules} already implies that the derivative of $K$ tends to zero.
A particular instance of $\log$-convex $K$ is the case where $\tau$ is the set of return times to zero of a one dimensional nearest-neighbor random walk on ${\ensuremath{\mathbb Z}} $. This is related to $\log$-convexity of the sequence of Catalan numbers (see \cite{cf:BF} for a paper on the subject).
\end{remark}
\subsection{The smoothing for polynomial tails}\label{smoothpol}
Let us explain briefly in this section why the strategy from \cite{cf:GT05} fails to give any results in the case of stretched exponential renewals (for more details the reader should refer to the original article). For simplicity we assume here that the environment is Gaussian and that $\beta=1$.
\medskip
The main idea in \cite{cf:GT05} is the following. Let $h=h_c(\beta)+u$ be fixed, and $N$ be chosen very large.
We look at a system at the critical point $h_c(\beta)$ (for which the free energy is zero):
in a typical segment of length $N$ the empirical mean of $\omega$ should be of order $0$ due to the law of large number ; however, with probability of order $\exp(- N u^2/2 )$ the empirical mean is larger than $u$. In that case, the system does not locally look critical and the partition function corresponding to the segment should be of order $e^{N \textsc{f}(\beta,h_c(\beta)+u)}$, if $N$ is chosen sufficiently large to avoid finite size effects.
\medskip
The distance between these segments of length $N$ which give an unusually "good" contribution to the partition function should be typically huge, that is, of order $\exp(u^2 N/2 )$, and thus the cost for making a huge jump between two consecutive good segments to avoid bad environment should be of order
$K(\exp(u^2 N/2))$. As the free-energy at criticality is zero, the strategy consisting in visiting all the "good" segments and avoiding all the bad ones should not give an exponentially large contribution to the partition function, hence the cost of making the large jump should completely compensate for the energy reward one gets when visiting a good segment.
For this reason one must have for sufficiently large $N$
\begin{equation}\label{kaboom}
K(\exp(u^2 N/2))e^{N \textsc{f}(\beta,h_c(\beta)+u)} < 1.
\end{equation}
In the case where $K$ has a power-law tail, this immediately yields a quadratic bound on the free energy. However, when $K$
has a lighter tail, \eqref{kaboom} fails to give any interesting information, as $K(\exp(u^2 N/2))$ decays super-exponentially.
\medskip
Some elements of this strategy can somehow be recycled (this is what is done in Section \ref{seclwb}) if one has some information about the behavior of finite volume systems (see Lemma \ref{finitevol}) . However, as will be seen, this fails to give a quadratic bound on the free-energy.
\subsection{Comparison with the case of renewals with exponential and sub-exponential tails}
An other instance of pinning model with absence of smoothing has been exhibited in \cite{cf:Aivy}: disordered pinning of transient renewals with exponential tails
(the case $K(n)=O(\exp(-n b))$ for some $b>0$).
However, let us mention that this is case quite special since when the tail of the renewal is exponential Remark \ref{rembc} is not valid anymore. On the contrary, the behavior of the system crucially depends on the contraint one imposes at the end point:
\begin{itemize}
\item The free-energy $\textsc{f}(\beta,h)$ defined by \eqref{freen}, which corresponds to a system constrained to be pinned, is negative for small values of $h$.
\item The free energy of the system with no constraints is obtained by considering the best of two strategies: either the walk will avoid the wall completely or it will try to pin the end point. The reward for this is equal to $\max(0,\textsc{f}(\beta,h))$, which is easily shown to have a first order transition in $h$.
\end{itemize}
Here the mechanism which triggers a first order phase transition is completely different: one has to perform an analysis of the local fluctuations of the environment to see whether or not the benefit of a good rare region is sufficient to compensate the cost of a large jump reaching it. An upper bound on the fluctuations is obtained via concentration.
To obtain a lower-bound, we choose to restrict to the Gaussian model for simplicity, but similar ideas could in principle be implemented by the use of tilting (like in \cite{cf:CdH}).
\section{Preliminaries}
\subsection{Notation}
The dependence in $\beta$ and $h$ will frequently be omitted to lighten the notation.
When $A$ is an event for $\tau$ we set
\begin{equation}
\label{eq:event}
Z^\omega_{N}(A):={\ensuremath{\mathbf E}} \left[e^{\sum_{n=1}^N(\beta\omega_n+h)\delta_n}\delta_N\mathbf{1}_A\right].
\end{equation}
For $k\in {\ensuremath{\mathbb N}} $, the shift operator $\theta^k$ acting on the sequence $\omega$ is defined by
\begin{equation}
\theta^k \omega_n:= \omega_{n+k}.
\end{equation}
For any couple of integers $a\le b$ we set
\begin{equation}\label{znab}
Z^{\omega}_{[a,b]}=e^{(\beta \omega_a+h)\mathbf{1}_{a> 0}} Z^{\theta^a \omega}_{b-a}.
\end{equation}
to be the partition function associated to the segment $[a,b]$ (with the convention that $Z^\omega_0=1$). Note that the environment at the starting point of the interval $a$ is taken into account only for $a>0$ (for technical reasons).
\medskip
For $\gep>0$ we define
\begin{equation}
\label{abepsilon}
\begin{split}
{\ensuremath{\mathcal A}} ^\gep&:=\{ \tau \ | \ \#( \tau\cap (0,N])\ | \le \gep N, N\in \tau \},\\
{\ensuremath{\mathcal B}} ^\gep&:=\{ \tau \ | \ \#( \tau\cap (0,N])\ | > \gep N, N\in \tau \},
\end{split}
\end{equation}
the set of renewals whose contact fraction is smaller, resp. larger, than $\gep$.
\subsection{Finite volume bounds for the free energy}
The following result allows to estimate the free-energy only knowing the value of $\frac{1}{N}{\ensuremath{\mathbb E}} \left[\log Z^\omega_N\right]$, for a given $N$.
\begin{lemma}\label{finitevol}
There exists a constant $c$ such that for every $N$, $\beta$ and $h$,
\begin{equation}\begin{split}
\frac{1}{N}{\ensuremath{\mathbb E}} \left[\log Z^\omega_N\right]& \le \textsc{f}(\beta,h),\\
\frac{1}{N}{\ensuremath{\mathbb E}} \left[\log Z^\omega_N\right]& \ge \textsc{f}(\beta,h)- \frac{N^{\zeta-1}}{1-2^{\zeta-1}}-\frac{2(\lambda(\beta)+h)_++c}{N}
\end{split}
\end{equation}
\end{lemma}
\begin{proof}
The first inequality is a consequence of following super-multiplicativity property
\begin{equation}
Z^{\omega}_{N+M}\ge Z^{\omega}_{N}\times Z^{\theta^N \omega}_{M}
\end{equation}
(see e.g.\ the proof of \cite[Proposition 4.2]{cf:GB}).
For the second one the proof is similar to \cite[Proposition 2.7]{cf:Alea}, one has
\begin{equation}\label{anelka}
Z^\omega_{2N}= {\ensuremath{\mathbf E}} \left[e^{\sum_{n=1}^N(\beta\omega_n+h)\delta_n}\delta_{N}\delta_{2N}\right]
+{\ensuremath{\mathbf E}} \left[e^{\sum_{n=1}^N(\beta\omega_n+h)\delta_n}(1-\delta_{N})\delta_{2N}\right].
\end{equation}
The first term is equal to
$Z^\omega_{N}Z^{\theta^N \omega}_{N}.$
For the second term, by comparing the weight of each $\tau$ to the one of $\tau\cup\{N\}$ one obtains
\begin{multline}\label{patagony}
{\ensuremath{\mathbf E}} \left[e^{\sum_{n=1}^N(\beta\omega_n+h)\delta_n}(1-\delta_{N})\delta_{2N}\right]\\
\le Z^\omega_{N}Z^{\theta^N \omega}_{N}e^{-\beta \omega_N-h}\max_{0\le a<N<b\le 2N} \frac{K(b-a)}{K(N-a)K(b-N)}\\
\le C e^{-\beta \omega_N-h} Z^\omega_{N}Z^{\theta^N \omega}_{N}\exp(N^{\zeta}),
\end{multline}
for some constant $C>1$.
The last line of \eqref{patagony} is obtained by observing that for any choice of $0\le a<N<b\le 2N$
$$(N-a)^\zeta +(b-N)^\zeta-(b-a)^\zeta\le N^{\zeta}.$$
Hence taking the $\log$ and expectation in \eqref{anelka} one has
\begin{equation}\begin{split}
\frac{1}{2N}{\ensuremath{\mathbb E}} \left[ \log Z^\omega_{2N}\right]&\le \frac{1}{N}{\ensuremath{\mathbb E}} \left[\log Z^\omega_{N}\right]+ \frac{1}{N} {\ensuremath{\mathbb E}} \left[\log \left( 1+Ce^{-\beta \omega_N-h} \exp(N^{\zeta})\right)\right]
\\
&\le \frac{1}{N}{\ensuremath{\mathbb E}} \left[ \log Z^\omega_{N}\right]+\frac{1}{N} \log \left( 1+e^{\lambda(-\beta)-h}C \exp(N^{\zeta})\right)\\
&\le \frac{1}{N}{\ensuremath{\mathbb E}} \left[ \log Z^\omega_{N}\right]+N^{\zeta-1}+\frac{1}{N} \log \left( 1+ C e^{\lambda(-\beta)-h}\right),\\
&\le \frac{1}{N}{\ensuremath{\mathbb E}} \left[ \log Z^\omega_{N}\right]+N^{\zeta-1}+\frac{1}{N} \left[ \log(2C) + (\lambda(-\beta)-h)_+ \right],\\
\end{split}\end{equation}
where the first inequality is simply Jensen's inequality.
Then we iterate the inequality and obtain
\begin{equation}
\textsc{f}(\beta,h)\le \frac{1}{N}{\ensuremath{\mathbb E}} \left[ \log Z^\omega_{N}\right]+\frac{N^{\zeta-1}}{1-2^{\zeta-1}}+\frac{2}{N} \left[ \log(2C) + (\lambda(-\beta)-h)_+ \right].
\end{equation}
\end{proof}
\subsection{The FKG inequality for $\log$-convex renewals}
For the proof of Theorem \ref{mainres3} (and only then), we need to use the fact that the presence of renewal point are positively correlated.
This is where we need the assumption on the $\log$ convexity of the function $K$.
\medskip
In this subsection $\tau$ denotes a subset of $\{1,\dots,N\}$ which contains $N$,
and with some abuse of notation ${\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N$ is considered to be a law on ${\ensuremath{\mathcal P}} (\{1,\dots,N\})$ (the set of subsets of $\{1,\dots,N\}$).
\medskip
Now let us introduce some definitions.
A function $f: {\ensuremath{\mathcal P}} (\{1,\dots,N\})\to {\ensuremath{\mathbb R}} $ is said to be increasing if
\begin{equation}
\forall \tau, \tau' \in {\ensuremath{\mathcal P}} (\{1,\dots,N\})\quad \tau\subset \tau' \Rightarrow f(\tau)\le f(\tau').
\end{equation}
Note that the following result was proved in \cite{cf:BW} for renewal processes in continuous time. Our proof is essentially similar and is based on the use of the
celebrated FKG criterion from \cite{cf:FKG} but we choose to include it for the sake of completeness.
\begin{proposition}
Assume that the function $K$ is $\log$-convex.
Then for all $\beta, \omega, h$ and $N$, the ${\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N$ satisfies the FKG inequality. Namely, for all increasing functions $f$ and $g$
\begin{equation}
{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_N[f(\tau) g(\tau)]\ge {\ensuremath{\mathbf E}} ^{\beta,h,\omega}_N[f(\tau)] {\ensuremath{\mathbf E}} ^{\beta,h,\omega}_N[g(\tau)].
\end{equation}
\end{proposition}
\begin{proof}
From \cite[Proposition 1]{cf:FKG}, it is sufficient to check that for any $\tau$ and $\tau'$ one has
\begin{equation}\label{ooo}
{\ensuremath{\mathbf P}} _N^{\beta,h,\omega}(\tau\cup \tau') {\ensuremath{\mathbf P}} _N^{\beta,h,\omega}(\tau\cap \tau') \ge {\ensuremath{\mathbf P}} _N^{\beta,h,\omega}(\tau) {\ensuremath{\mathbf P}} _N^{\beta,h,\omega}(\tau').
\end{equation}
For $\sigma\subset \{0,\dots,N\}$ whose elements are $\sigma_0=0<\sigma_1<\dots<\sigma_m=N$, we set
$$K(\sigma)=\prod_{i=1}^m K(\sigma_i-\sigma_{i-1}).$$
The reader can check that after simplification \eqref{ooo} is equivalent to
\begin{equation}\label{toto}
K(\tau\cup \tau')K(\tau\cap \tau') \ge K(\tau) K(\tau').
\end{equation}
This inequality is obviously true when $\tau'\subset \tau$. Then we proceed by induction and it is sufficient to
check that if $a\notin \tau\cup \tau' $ and \eqref{toto} holds for $\tau$ and $\tau'$, then it holds for
$\tau$ and $\tau'\cup\{a\}$. To this purpose we only need to check that for any $\tau$, $\tau'$ and $a\notin \tau\cup \tau' $ we have
\begin{equation}\label{hopla}
\frac{K(\tau\cup \tau' \cup\{a\})}{K(\tau\cup \tau')}\ge \frac{K(\tau' \cup\{a\})}{K(\tau')}.
\end{equation}
Let us set
\begin{equation}
\begin{split}
\alpha_1:=\inf\{ x < a \ | \ x\in \tau\cup \tau' \}, \quad & \beta_1:=\inf\{ x > a \ | \ x\in \tau\cup \tau' \},\\
\alpha_2:=\inf\{ x < a \ | \ x\in \tau' \}, \quad & \beta_2:=\inf\{ x > a \ | \ x\in \tau' \}.
\end{split}
\end{equation}
We remark that $$\alpha_2\le \alpha_1<a<\beta_1\le \beta_2.$$
The inequality \eqref{hopla} is equivalent to
\begin{equation}\label{totti}
\frac{K(\beta_1-a)K(a-\alpha_1)}{K(\beta_1-\alpha_1)}\ge \frac{K(\beta_2-a)K(a-\alpha_2)}{K(\beta_2-\alpha_2)}.
\end{equation}
By convexity of $\log K$ the function
\begin{equation}
(\alpha,\beta)\mapsto \frac{K(\beta-a)K(a-\alpha)}{K(\beta-\alpha)}
\end{equation} is non-increasing in $\beta$ and non-decreasing in $\alpha$ on the set $\{ (\alpha,\beta) \ | \ \alpha<a<\beta \}$ . Thus \eqref{totti} holds.
\end{proof}
\section{Proof of Theorem \ref{mainres}}
\subsection{Decomposition of the proof}
The key point consists in proving the following upper bound on $Z_N^\omega({\ensuremath{\mathcal A}} ^\gep)$ (recall \eqref{abepsilon} and \eqref{eq:Znh}).
\begin{proposition}\label{keypoint}
There exist positive constants $\gep_0$ and $C$ such that for all $\gep\le \gep_0$
we have for all $h\le 1$ and $\beta>0$, , almost surely for all $N$ sufficiently large,
\begin{equation}\label{orange}
\frac{1}{N} \log Z^\omega_N({\ensuremath{\mathcal A}} ^\gep)\le \frac{1}{2}\textsc{f}(h,\beta) + \max_{l\ge \gep^{-1}} \left( C\beta \sqrt{\frac{\gep \log l}{l}}-\frac{1}{4}l^{\zeta-1} \right).
\end{equation}
\end{proposition}
The restriction $h\le 1$ is chosen for convenience but does not convey any particular significance ($h<c$ for some $c>0$ would be just as good).
The proof of this statement is postponed to Section \ref{quatpoind}.
Now, we observe that if $\gep$ is chosen to be larger than the asymptotic contact fraction $\partial_h \textsc{f}(\beta,h)$, the l.h.s.\ of \eqref{orange} converges to the the free-energy.
\begin{lemma}\label{fractio}
For every $h>h_c(\beta)$ when $\gep>\partial_h\textsc{f}(\beta,h)$ one has.
\begin{equation}\label{limimi}
\liminf_{N\to \infty} {\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N\left[{\ensuremath{\mathcal A}} ^\gep\right]>0.
\end{equation}
As a consequence
\begin{equation}
\limsup_{N\to \infty} \frac{1}{N} \log Z^\omega_N({\ensuremath{\mathcal A}} ^\gep)=\textsc{f}(\beta,h).
\end{equation}
\end{lemma}
\begin{remark}
Without much more efforts, one can even prove in fact that the limit in \eqref{limimi} is equal to one, but this is not necessary for our purpose.
\end{remark}
The idea to prove Theorem \ref{mainres} is to use \eqref{orange} where $\gep$ is replaced by $2\partial_h \textsc{f}(\beta,h)$
and $\frac{1}{N} \log Z_N^\omega({\ensuremath{\mathcal A}} ^\gep)$ is replaced by its limit: $\textsc{f}(\beta,h)$. This gives a differential inequality in $h$ which once integrated gives the claimed bounds on the free energy. Details follow at the end of the Section.
\begin{proof}[Proof of Lemma \ref{fractio}]
For simplicity (and with no loss in generality) assume that $\gep=(1+\delta) \partial_h\textsc{f}(\beta,h)$ for some $\delta<1$.
By \eqref{gro}, for $N$ sufficiently large
\begin{equation}
\frac{1}{N}{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_N\left[\sum_{n=1}^N \delta_n\right]\le (1+\delta)(1-\delta/2) \partial_h \textsc{f}(\beta,h)=(1-\delta/2) \gep.
\end{equation}
As
\begin{equation}
\frac{1}{N}{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_N\left[\sum_{n=1}^N \delta_n\right]\ge \gep {\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N \left[{\ensuremath{\mathcal B}} ^\gep\right],
\end{equation}
this implies
\begin{equation}
{\ensuremath{\mathbf P}} ^{\beta,h,\omega}_N\left[{\ensuremath{\mathcal B}} ^\gep\right]\le 1-\delta/2.
\end{equation}
\end{proof}
\begin{proof}[Proof of Theorem \ref{mainres}]
Let $\gep_0$ be such that Proposition \ref{keypoint} holds (in the case $\zeta>1/2$, we will also require $\gep_0$ to satisfy another condition).
If $h\le 1$ is such that
\begin{equation}\label{defeph}
\gep_h:=2\partial_h\textsc{f}(\beta,h)\le \gep_0,
\end{equation}
we have
\begin{equation}\label{cramook}
\textsc{f}(\beta,h) \le \max_{l\ge \gep_h^{-1}}\left( 2C\beta\sqrt{\frac{\gep_h \log l}{l}}- \frac{1}{2}l^{\zeta-1} \right).
\end{equation}
Let us start with the case $\zeta>1/2$. By contradiction, let us assume that,
\begin{equation}\label{mangepasdpin}
\lim_{h\to h_c(\beta)+} \partial_h \textsc{f}(\beta,h)< \frac{1}{2}\gep_0\min(1,\beta^{-2}).
\end{equation}
From a standard convexity argument (see \cite[Proposition 5.1]{cf:GB}) one has $h_c(\beta)\le 0$. Thus we can find
$h\in (h_c(\beta), 1]$ such that
$$\gep_h\le \gep_0\min(1,\beta^{-2}).$$
Then for this value of $h$ the right-hand side of \eqref{cramook} is smaller than
$$ \sup_{l\ge \gep_0^{-1}}\left( 2C\sqrt{\frac{\gep_0 \log l}{l}}- \frac{1}{2}l^{\zeta-1} \right)$$
which is equal to zero if $\gep_0$ has been chosen sufficiently small. Hence we obtain a contradiction as $\textsc{f}(\beta,h)>0$.
\medskip
Let us move to the case $\zeta\le1/2$. We can assume that
\begin{equation}\label{mangepasdpin2}
\lim_{h\to h_c(\beta)+} \partial_h \textsc{f}(\beta,h)=0
\end{equation}
as if not, there is nothing to prove.
\medskip
For $h$ sufficiently close to the critical point we hence have $\gep_h\le \gep_0$ (and $h\le 1$)
and hence
Equation \eqref{cramook} holds.
Computing the maximum in the r.h.s.\ of \eqref{cramook} we obtain
\begin{equation}\label{ineqps}
\textsc{f}(\beta,h)\le \begin{cases} C\left( \beta^2 \gep_h |\log \gep_h |\right)^{\frac{1-\zeta}{1-2\zeta}} &\text{ for } \zeta<1/2 \\
\exp(- c(\beta^2 \gep_h)^{-1}) &\text{ for } \zeta= 1/2.
\end{cases}
\end{equation}
When $\zeta<1/2$, we note that the inverse of the function
$$f: x \to C \left( \beta^2 x | \log x |\right) ^{\frac{1-\zeta}{1-2\zeta}} $$
(which is increasing near zero)
satisfies
\begin{equation}
f^{-1}(y)\stackrel{y\to 0} {\sim} C ' \beta^{-2} y^{\frac{1-2\zeta}{1-\zeta}} | \log y |^{-1}.
\end{equation}
Hence composing the inequality \eqref{ineqps} with $f^{-1}$ and replacing $\gep_h$ by its value \eqref{defeph}, we obtain that in the vicinity of $h_c(\beta)_+$ we have
\begin{equation}
\partial_h \textsc{f} \ge c \beta^{-2} \textsc{f}^{\frac{1-2\zeta}{1-\zeta}} | \log \textsc{f} |^{-1},
\end{equation}
and hence
\begin{equation}\label{tobeintegrated}
\textsc{f}^{\frac{2\zeta-1}{1-\zeta}} |\log \textsc{f}| \partial_h \textsc{f} \ge c \beta^{-2}.
\end{equation}
It is easy to check that this inequality is valid also in the case $\zeta=1/2$.
Now at the cost of modifying the constant $c$ (and the neighborhood of $h_c(\beta)$ if necessary)
the inequality implies
\begin{equation}\label{tobeintegrated2}
\partial_h \left[ \textsc{f}^{\frac{\zeta}{1-\zeta}} |\log \textsc{f}| \right] \ge c \beta^{-2},
\end{equation}
which once integrated implies
\begin{equation}
\textsc{f}(\beta,h_c(\beta)+u)^{\frac{\zeta}{1-\zeta}} |\log \textsc{f}(\beta,h_c(\beta+u)|\ge c\beta^{-2}u.
\end{equation}
Composing the inequality with the inverse of the function $x\mapsto x^{\frac{\zeta}{1-\zeta}} |\log x|$ near zero gives the desired result.
Integrating the above inequality between $h_c(\beta)$ and $h$ yields the result.
\end{proof}
\subsection{Proof of Proposition \ref{keypoint}}\label{quatpoind}
A key tool in the proof is the following concentration inequality.
\begin{lemma}\label{concentration}
When Assumption \ref{logsob} holds then
for any event $A\subset {\ensuremath{\mathcal A}} ^\gep$
\begin{equation}
{\ensuremath{\mathbb P}} \left[\log Z^{\omega}_N(A)-{\ensuremath{\mathbb E}} [\log Z^{\omega}_N(A)]\ge t\right]\le C_1 \exp\left(-\frac{t^2}{C_2\beta^2 N\gep}\right).
\end{equation}
\end{lemma}
\begin{proof}
For any pair of environment $\omega$ and $\omega'$ one has
\begin{equation}
\left | \log \frac{ Z^{\omega}_N(A)}{Z^{\omega'}_N(A)} \right |\le \beta \max_{\big\{ \tau \subset [0,N] \ | \ |\tau \cap [0,N]|\le \gep N \big\}} \sum_{x\in \tau}
|\omega_x-\omega'_x|\le \beta \sqrt{\gep N} \sqrt{\sum_{x=1}^N \omega^2_x}.
\end{equation}
Hence the Lipshitz norm of
$$\omega\mapsto \log Z^{\omega}_N(A)$$ is smaller than $\beta\sqrt{\gep N}$. It is also a convex function, thus the results follows from Assumption \ref{logsob}.
\end{proof}
Given $\tau \in {\ensuremath{\mathcal A}} ^\gep$, we define ${\ensuremath{\mathcal L}} (\tau)$ the set of indices corresponding to the renewal jumps of length larger than $(2\gep)^{-1}$, and we denote by $L(\tau)$ the cardinal of ${\ensuremath{\mathcal L}} (\tau)$.
\begin{equation}
\begin{split}
{\ensuremath{\mathcal L}} (\tau)&:=\{ n \ | \ \tau_{n}\le N, (\tau_{n}-\tau_{n-1})\ge (2\gep)^{-1} \},\\
L(\tau)&:=\#{\ensuremath{\mathcal L}} (\tau).
\end{split}\end{equation}
We also set $l(\tau)=N/L(\tau)$.
Due to the definition of ${\ensuremath{\mathcal A}} ^\gep$ one has
\begin{equation}
\forall \tau \in {\ensuremath{\mathcal A}} ^\gep, \quad \sum_{n\in {\ensuremath{\mathcal L}} (\tau)} (\tau_{n}-\tau_{n-1})\ge \frac{N}{2}.
\end{equation}
This means in particular than $l$ is roughly the mean length of $(\tau_{n}-\tau_{n-1})_{n\in {\ensuremath{\mathcal L}} (\tau)}$ (up to a factor $2$).
For a fixed $L\in {\ensuremath{\mathbb N}} $, $L\le \gep N$, we set
\begin{multline}
{\ensuremath{\mathcal T}} (L):= \left\{ ({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )\in ([0,N]\cap {\ensuremath{\mathbb Z}} )^{2L} \ \big| \ \forall i\in [1,L], \ t'_i\ge t_{i-1}, \ t_{i}\ge t'_i+(2\gep)^{-1} \right\}\\
\cap \left\{ \sum_{i=1}^L (t_i-t'_i)\ge N/2\right\},
\end{multline}
which is the set of possible locations for $(\tau_{n-1},\tau_n)_{n\in {\ensuremath{\mathcal L}} (\tau)}$.
For $({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )\in {\ensuremath{\mathcal T}} (L)$ we set
\begin{equation}
A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )}:=\left\{ \tau \ \big| \ \{ (\tau_{n-1},\tau_n) \}_{n\in {\ensuremath{\mathcal L}} (\tau)}= \{ (t'_i ,t_i) \}_{i=1}^L \ \right\}\cap {\ensuremath{\mathcal A}} ^{\gep}.
\end{equation}
It is the subset of ${\ensuremath{\mathcal A}} ^\gep$ for which the jumps of $\tau$ which are longer than $(2\gep)^{-1}$ exactly span the segments
$(t'_i ,t_i)_{i=1}^L$ (see also Figure \ref{attprim}).
\begin{figure}[hlt]
\epsfxsize =11.5 cm
\begin{center}
\psfrag{O}{$0$}
\psfrag{N}{$N$}
\psfrag{v1}{$t'_1$}
\psfrag{v2}{$t'_2$}
\psfrag{t1}{$t_1$}
\psfrag{t3}{$t_3$}
\psfrag{t2=v3}{$t_2=t'_3$}
\epsfbox{largjumps.eps}
\end{center}
\caption{\label{attprim}
A schematic representation of a set $({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )\in {\ensuremath{\mathcal T}} (3)$, and a renewal in $\tau\in A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )}$ (in green).
The total number of jumps must be smaller than $\gep N$ and in yellow regions, the jumps of $\tau$ must be shorter than $(2\gep)^{-1}$.
As a consequence of these two conditions the total length of the yellow regions is smaller than $N/2$.}
\end{figure}
We have
\begin{equation}
Z_N^\omega({\ensuremath{\mathcal A}} ^\gep)= \sum_{L=1}^{\gep N} \sum_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )\in {\ensuremath{\mathcal T}} (L)} Z_N^\omega (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )}).
\end{equation}
In particular
\begin{equation}\label{cramnik}
\log Z_N^\omega({\ensuremath{\mathcal A}} ^\gep) \le \log N + \max_{L\in\{1,\dots,\gep N\}} \left[ \log (\# {\ensuremath{\mathcal T}} (L))+ \max_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} ) \in {\ensuremath{\mathcal T}} (L) } \log Z_N^\omega (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )} )\right].
\end{equation}
The idea is then to use Lemma \ref{concentration} to find a good bound on the l.h.s.
A first easy step is to get an estimate on the cardinal of ${\ensuremath{\mathcal T}} (L)$. Recall that here and in what follows $l:=N/L$.
\begin{lemma}\label{one}
There exists a $C$ such that for all $\gep\le 1/4$ for all $N$ sufficiently large
\begin{equation}
\# {\ensuremath{\mathcal T}} (L)\le C \exp( 2 L \log l ).
\end{equation}
\end{lemma}
\begin{proof}
The set $\{ t_i\}_{i=1}^L\cup \{ t'_i+1\}_{i=1}^L$ is a subset of
$\{1,\dots, N\}$ with $2L$ elements. Hence
\begin{equation}
\# {\ensuremath{\mathcal T}} (L)\le \binom{N}{2L}\le C \exp( 2 L \log l ),
\end{equation}
where the last inequality just comes from Stirling formula.
\end{proof}
To use Lemma \ref{concentration} efficiently, we must also know about the expected value of $\log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})$
\begin{lemma}\label{shanti}
For any $({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} ) \in {\ensuremath{\mathcal T}} (L)$, one has, for $\gep$ sufficiently small (depending only on $K$)
\begin{equation}
\frac{1}{N}{\ensuremath{\mathbb E}} [ \log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})] \le \frac{1}{2}\left( \textsc{f}(\beta,h) + l^{-1} - l^{\zeta-1}\right).
\end{equation}
\end{lemma}
\begin{proof}
One has (recall \eqref{znab})
\begin{equation}
Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})\le \left[ \prod_{i=1}^L Z_{[t_{i-1},t'_{i}]}K(t_i-t'_i)\right] Z_{[t_L,N]},
\end{equation}
where we take the convention $t_0=0$.
Hence
\begin{equation} \label{sketletor}
{\ensuremath{\mathbb E}} \left[\log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )}\right]\le \sum_{i=1}^L {\ensuremath{\mathbb E}} \left[\log Z_{[t_{i-1},t'_{i}]} \right] + {\ensuremath{\mathbb E}} \left[ \log Z_{[t_L,N]}\right]
+\sum_{i=1}^L \log K(t_i-t'_i).
\end{equation}
One has from Lemma \ref{finitevol}
for all $i>1$
$${\ensuremath{\mathbb E}} Z_{[t_{i-1},t'_{i}]}\le (t'_{i}-t_{i-1})\textsc{f}(\beta, h) + h,$$
the extra $h$ term is there because the definition of $Z_[a,b]$ \eqref{defznab} takes into-account the environment at the starting point (
and the fact that $h\le 1$, it follows that
\begin{multline}
\sum_{i=1}^L {\ensuremath{\mathbb E}} \left[\log Z_{[t_{i-1},t'_{i}] }\right] + {\ensuremath{\mathbb E}} \left[ \log Z_{[t_L,N]}\right] \\
\le \left(\sum_{i=1}^L(t'_i-t_{i-1})+ (N-t_L)\right)\textsc{f}+ L h
\le N \textsc{f}/2+ L.
\end{multline}
is not exactly the partition function (
Regarding the last term in \eqref{sketletor}, using Jensen's inequality for the function $x\mapsto x^{\zeta}$, we have, choosing $\delta>0$ sufficiently small, for $\gep$ sufficiently small
\begin{equation}
-\sum_{i=1}^L \log K(t_i-t'_i)\ge (1-\delta) \sum_{i=1}^L (t_i-t'_i)^{\zeta}\ge (1-\delta) 2^{-\zeta} L l^{\zeta}\ge \frac{1}{2}L l^{\zeta},
\end{equation}
which ends the proof.
\end{proof}
\begin{lemma}\label{ohm}
There exists a constant $C$ such that for $N$ sufficiently large,
for all $L\in \{1,\dots,\gep N\}$,
\begin{equation}
{\ensuremath{\mathbb P}} \left( \max_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} ) \in {\ensuremath{\mathcal T}} (L)} \Big( \log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})- {\ensuremath{\mathbb E}} [ \log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})]\Big)\ge C\beta N\sqrt{\frac{\gep \log l}{l}} \right)\le
\frac{1}{N^3}.
\end{equation}
\end{lemma}
\begin{proof}
From Lemma \ref{concentration} and a standard union bounds one has for any $u$
\begin{equation}
{\ensuremath{\mathbb P}} \left( \max_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} ) \in {\ensuremath{\mathcal T}} (L)} \left( \log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})- {\ensuremath{\mathbb E}} [ \log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})]\right) \ge u\right)\le (\#{\ensuremath{\mathcal T}} )C_1 \exp\left(-\frac{u^2}{C_2\beta^2 N\gep}\right)
\end{equation}
Using Lemma \ref{one} for the value of $u:=C\beta N\sqrt{\frac{\gep \log l}{l}}$ one can conclude provided that $C$ is chosen sufficiently large.
\end{proof}
\begin{proof}[Proof of Proposition \ref{keypoint}]
Using Lemma \ref{ohm} and Lemma \ref{shanti},
one has almost surely for all large $N$, for all $L\le \gep N$
\begin{equation}
\frac{1}{N}\max_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} ) \in {\ensuremath{\mathcal T}} (L)} \log Z^{\omega}_N (A_{({\ensuremath{\mathbf t}} ',{\ensuremath{\mathbf t}} )})\le \frac{1}{2}\textsc{f}(\beta,h)+l^{-1}+C\beta \sqrt{\frac{\gep \log l}{l}}-\frac{1}{2} l^{\zeta-1}.
\end{equation}
Combining this with \eqref{cramnik} we obtain
\begin{equation}
\frac 1 N\log Z_N^\omega({\ensuremath{\mathcal A}} ^\gep)\le \frac{\log N} N + \max_{l\ge \gep^{-1}} \left( \frac{\log \# {\ensuremath{\mathcal T}} (L)}{N}+\frac{1}{2}\textsc{f}(\beta,h)+ l^{-1}+ C\beta \sqrt{\frac{\gep \log l}{l}}-\frac{1}{2} l^{\zeta-1}\right).
\end{equation}
The terms $\frac{\log N} N$ and $\frac{\log \# {\ensuremath{\mathcal T}} (L)}{N}$ can be neglected if $l$ is sufficiently large (i.e. $\gep$ is sufficiently small) and $l^{\zeta-1}/2$ is replaced by $l^{\zeta-1}/4$.
\end{proof}
\section{Proof of Theorem \ref{mainres2}: rounding for $\zeta< 1/2$}\label{seclwb}
The idea to find an upper bound on the free-energy is close to the one in \cite{cf:GT05}. The main difference is that here, we must combine the argument with the finite volume criterion given by Lemma \ref{finitevol} to get a result.
We use the fact that $\omega$ is Gaussian in the following way:
\begin{lemma}\label{gaussianlemma}
For any $N$ if $\omega$ are IID Gaussian variables then the
sequence
$$\left(\omega_x-\frac{1}{N}\sum_{n=1}^N \omega_n\right)_{x=1}^N$$
is independent of
$\sum_{n=1}^N \omega_n$.
\end{lemma}
With this observation, we see that changing the value of $h$ by an amount $\delta$ is in fact equivalent to changing the empirical mean of the $\omega$ by an amount $\delta \beta^{-1}$.
\medskip
In a first step we try to control the expectation of the free-energy for a typical value of $\sum_{n=1}^N \omega_n$.
\begin{proposition}\label{bound}
There exists a constant $C$ such that
for all $N$ sufficiently large and all $u$,
\begin{equation}\label{dabound}
{\ensuremath{\mathbb E}} \left[\log Z^{\beta,h_c(\beta),\omega}_N \ \big| \ \sum_{n=1}^N \omega_n\ge u \sqrt{N} \right]
\le C N^{\zeta}(1+|u|^{\zeta})e^{\zeta u^2/2}+\beta^2.
\end{equation}
\end{proposition}
This will be done using the finite volume criterion of Lemma \ref{finitevol}: if \eqref{dabound} does not hold, one can find a strategy which gives a positive free-energy for $h=h_c(\beta)$ and hence yields a contradiction.
Then the idea is to integrate this bound over all values of $u$ to obtain a bound for ${\ensuremath{\mathbb E}} \left[\log Z^{\beta,h,\omega}_N\right]$.
Of course the bound will be a good one only if $N$ is wisely chosen. We can finally conclude using the finite volume criterion Lemma \ref{finitevol}.
\begin{proof}[Proof of Theorem \ref{mainres2}]
Now for $h=h_c(\beta)+v$ one sets $N:=\beta^2 v^{-2}$ (assuming that we have chosen $v$ such that $N$ is an integer).
One has
\begin{equation}\begin{split}
{\ensuremath{\mathbb E}} \left[ \log Z^{\beta,h,\omega}_N\right]&= \int \frac{1}{\sqrt{2\pi}}\exp\left(-u^2/2\right){\ensuremath{\mathbb E}} \left[\log Z^{\beta,h,\omega}_N \ | \ \sum_{n=1}^N \omega_n= u \sqrt{N} \right]\dd u\\
&=\int \frac{1}{\sqrt{2\pi}}\exp\left(-\frac{(u-\beta^{-1} v\sqrt{N})^2}{2}\right){\ensuremath{\mathbb E}} \left[\log Z^{\beta,h_c(\beta),\omega}_N \ | \ \sum_{n=1}^N \omega_n= u \sqrt{N} \right]\dd u.
\end{split}
\end{equation}
Using Proposition \ref{bound} we have the following inequality provided that $v$ is sufficiently small (in which case the $\beta^2$ can be neglected)
\begin{equation}
{\ensuremath{\mathbb E}} \left[ \log Z^{\beta,h,\omega}_N\right]\le C N^{\zeta} \int \frac{1}{\sqrt{2\pi}}(1+|u|^{\zeta}) \exp\left(\frac{\zeta u^2-(u-1)^2}{2}\right)\dd u\le C' N^{\zeta}.
\end{equation}
Hence, using Lemma \ref{finitevol}, we obtain
\begin{equation}
\textsc{f}(\beta,h)\le C'' N^{\zeta-1}= C'' (v\beta^{-1})^{2(1-\zeta)}.
\end{equation}
\end{proof}
\begin{proof}[Proof of Proposition \ref{bound}]
One can assume $u\ge 1$ without loss of generality.
Set $$M:= u\exp\left(u^2/2\right).$$
Let $X_0$ be the smallest integer such that
\begin{equation}
\sum_{n=X_0 N+1}^{(X_0+1)N} \omega_x\ge u \sqrt{N}.
\end{equation}
Then we obtain a lower bound on $Z_{NM}$ by deciding to visit the stretch $[X_0N,(X_0+1)N]$ if $X_0\le M- 2$
and to do only a long excursion in the other case (recall \eqref{znab}) (see Figure \ref{factorx}):
\begin{figure}[hlt]
\epsfxsize =11.5 cm
\begin{center}
\psfrag{O}{$0$}
\psfrag{NM}{$NM$}
\psfrag{X0N}{\tiny {$X_0N$}}
\psfrag{X01N}{\tiny {$(X_0+1)N$}}
\epsfbox{factorx.eps}
\end{center}
\caption{\label{factorx}
Here we present our strategy to obtain a lower bound on the partition function $Z_{NM}$. The yellow segments are those which are such that the empirical mean of $\omega$ is larger than $u N^{-1/2}$.
By definition $[X_0N,(X_0+1)N]$ is the first of these segments. We allow pinning only on the segment $[X_0N,(X_0+1)N]$ and only if $X_0\le M-2$. }
\end{figure}
\begin{equation}
Z^\omega_{MN}\ge \begin{cases} K(X_0N) Z^\omega_{[X_0N,(X_0+1)N]}K((M-X_0+1)N)e^{\beta \omega_{NM}+h_c(\beta)}, &\text{ if } X_0\le (M-2), \\
K(MN)e^{\beta \omega_{NM}+h_c(\beta)} \quad &\text{ if } X_0\ge M-2.
\end{cases}
\end{equation}
Taking the expectation one obtains, by translation invariance
\begin{multline}
{\ensuremath{\mathbb E}} \left[\log \left( K(X_0N) \log Z^\omega_{[X_0N,(X_0+1)N]}K((M-X_0+1)N)\right) \ | \ X_0\le (M-2) \right]\\
\ge -2(MN)^{\zeta}+h_c(\beta)+{\ensuremath{\mathbb E}} \left[\log Z^{\beta,h_c(\beta),\omega}_N \ | \ \sum_{n=1}^N \omega_x\ge u \sqrt{N} \right].
\end{multline}
We also have (as $\omega_{MN}$ is independent of the event $\{X_0\le M-2\}$ its conditional mean is zero)
\begin{equation}
{\ensuremath{\mathbb E}} \left[ \log \left[K(MN)e^{\beta \omega_{NM}+h_c(\beta)}\right] \ | \ X_0\ge M-2 \right]= \log K(MN)+h_c(\beta).
\end{equation}
And hence (recall that $h_c(\beta)\ge -\beta^2/2$ for Gaussian environments),
\begin{equation}
{\ensuremath{\mathbb E}} [\log Z^\omega_{MN}]\ge {\ensuremath{\mathbb E}} \left[\log Z^{\beta,h_c(\beta),\omega}_N \ | \ \sum_{n=1}^N \omega_x\ge u \sqrt{N} \right] {\ensuremath{\mathbb P}} \left[ X_0\le M-2 \right]-2 (MN)^{\zeta}-\beta^2.
\end{equation}
By standard estimates on Gaussian tails, there exists a constant $c>0$ such that
$$\forall u >1, \quad {\ensuremath{\mathbb P}} \left[ \sum_{n=1}^N \omega_x\ge u \sqrt{N} \right]\ge \frac c u e^{-u^2/2},$$ and hence, using the definition of $M$
we have
$${\ensuremath{\mathbb P}} \left[ X_0\le M-2 \right]>c',$$
for some positive constant $c'$.
This implies (recall Lemma \ref{finitevol} and that $\textsc{f}(\beta,h_c(\beta))=0$) that there exists $c''>0$ such that
\begin{equation}
0\ge {\ensuremath{\mathbb E}} [\log Z^\omega_{MN}]\ge c' \left( {\ensuremath{\mathbb E}} \left[\log Z^{\beta,h_c(\beta),\omega}_N \ | \ \sum_{n=1}^N \omega_x\ge u \sqrt{N} \right]\right) -c'' u^{\zeta} e^{\zeta u^2/2} N^\zeta-\beta^2.
\end{equation}
The above inequality is in fact only valid if one assumes that
$$
{\ensuremath{\mathbb E}} \left[\log Z^{\beta,h_c(\beta),\omega}_N \ | \ \sum_{n=1}^N \omega_x\ge u \sqrt{N} \right]\ge 0,$$
but if this is not the case there is nothing to prove.
\end{proof}
\section{Proof of Theorem \ref{mainres3}: rounding for $\zeta= 1/2$}
The case for $\zeta=1/2$ is a bit more complicated.
Assume that
\begin{equation}\label{defc0}
\lim_{h\to h_c(\beta)+} \partial_h\textsc{f}(\beta,h)=c_0>0,
\end{equation}
and let us derive a contradiction.
Fist, we prove that the contact fraction at the critical point, if well defined, cannot be equal to $c_0$ as there is always a positive probability for
the polymer to have a very small contact fraction.
\begin{lemma}\label{criticool}
The following three statements hold
\begin{itemize}
\item [(i)]
For all $\gep>0$, one has
\begin{equation}
\limsup_{N\to \infty} {\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^\gep \right) \right]<1.
\end{equation}
\item [(ii)]For any $u>c_0$ one has
\begin{equation}
\lim_{N\to \infty} {\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^u \right) \right]=0.
\end{equation}
\item [(iii)]One has
\begin{equation}
\limsup_{N\to \infty} \frac 1 N {\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h_c(\beta),\omega}_N\left(\sum_{n=1}^N\delta_N \right) \right]<c_0.
\end{equation}
\end{itemize}
\end{lemma}
\begin{proof}
Point $(iii)$ is a simple consequence of the two first point as
\begin{equation}
\frac 1 N {\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h_c(\beta),\omega}_N\left(\sum_{n=1}^N\delta_N \right) \right]=\frac 1 N\int_0^1 {\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^u \right) \right]\dd u.
\end{equation}
Point $(ii)$ is rather easy to prove:
Assume that for $u>c_0$ and for some $\delta>0$ one has
\begin{equation}
{\ensuremath{\mathbb P}} \left[{\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^u \right)>\delta\right]>\delta,
\end{equation}
for infinitely many $N$.
We note that if
$${\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^u \right)>\delta,$$ then
\begin{equation}
Z^{\beta,h_c(\beta),\omega}_N({\ensuremath{\mathcal B}} ^u)\ge \delta Z^{\beta,h_c(\beta),\omega}_N \ge \delta K (N)e^{\beta\omega_N+h_c(\beta)},
\end{equation}
where the last inequality is just obtained by considering renewal trajectories with only one contact.
Hence, for every $h>h_c(\beta)$ we have
\begin{equation}
Z^{\beta,h,\omega}_N\ge Z^{\beta,h,\omega}_N({\ensuremath{\mathcal B}} ^u)\ge \delta e^{Nu(h-h_c)} K(N)e^{\beta\omega_N+h_c(\beta)}.
\end{equation}
This implies (as we know that the limit exists and is non-random) that for every $h>h_c(\beta)$
\begin{equation}
\lim_{N\to \infty} \frac{1}{N}\log Z^{\beta,h,\omega}_N \ge u(h-h_c(\beta))
\end{equation}
which contradicts assumption \eqref{defc0} for small $h$.
\medskip
To prove $(i)$ let us assume that
\begin{equation}\label{opc}
\lim_{N\to \infty} {\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^\gep \right) \right]=1,
\end{equation}
(or that it occurs along a subsequence) and derive a contradiction from it.
Set
\begin{equation}
f_N(u):={\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^\gep \right) \ | \ \sum_{x=1}^{N-1} \omega_x= u\sqrt{N-1} \right].
\end{equation}
We have
\begin{equation}
{\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^\gep \right) \right]=\int\frac{1}{\sqrt{2\pi}}\exp\left(-u^2/2\right)f_N(u)\dd u.
\end{equation}
As $f_N(u)$ is an increasing function of $u$ this implies that for all $u\in {\ensuremath{\mathbb R}} $
\begin{equation}
\lim_{N\to\infty} f_N(u)=1.
\end{equation}
Fix $u=-10 \gep^{-1}$ and let $N$ be sufficiently large so that $f_N(u)\ge 3/4$.
Then necessarily
\begin{equation}\label{sixquatorz}
{\ensuremath{\mathbb P}} \left( {\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^\gep \right)\ge 1/2 \ | \ \sum_{x=1}^{N-1} \omega_x= u\sqrt{N-1} \right)\ge 1/2.
\end{equation}
Note that ${\ensuremath{\mathbf P}} ^{\beta,h_c(\beta),\omega}_N\left({\ensuremath{\mathcal B}} ^\gep \right)\ge 1/2$ implies in particular that
$$ Z^{\beta,h_c(\beta),\omega}_N({\ensuremath{\mathcal B}} ^\gep)\ge Z^{\beta,h_c(\beta),\omega}_N(({\ensuremath{\mathcal B}} ^\gep)^c)\ge K(N)e^{\beta\omega_N+h_c(\beta)}.$$
And hence \eqref{sixquatorz}
\begin{equation}
{\ensuremath{\mathbb P}} \left( Z^{\beta,h_c(\beta),\omega}_N({\ensuremath{\mathcal B}} ^\gep)\ge K(N)e^{\beta \omega_N+h_c(\beta)} \ | \ \sum_{x=1}^{N-1} \omega_x= u\sqrt{N-1} \right)\ge 1/2.
\end{equation}
From Lemma \ref{gaussianlemma}, replacing $u$ by $v$ in the conditioning is equivalent to replacing $\omega_n$ by $\omega_n+(v-u)(N-1)^{-1/2}$
for $n\in \{1,\dots,N-1\}$.
Hence for $v\ge u$ we have
\begin{equation}\label{decompo}
{\ensuremath{\mathbb P}} \left( Z^{\beta,h_c(\beta),\omega}_N({\ensuremath{\mathcal B}} ^\gep)\ge K(N) e^{\gep(v-u) \sqrt{N-1}+\beta\omega_N+h_c(\beta)} \ | \ \sum_{x=1}^{N-1} \omega_x= v\sqrt{N-1} \right)\ge 1/2.
\end{equation}
This implies that for any $v$ (this is obvious for $v\le u$)
\begin{equation}
{\ensuremath{\mathbb P}} \left( Z^{\beta,h_c(\beta),\omega}_N\ge K(N) e^{\gep(v-u) \sqrt{N-1}+\beta\omega_N+h_c(\beta)} \ | \ \sum_{x=1}^{N-1} \omega_x= v\sqrt{N-1} \right)\ge 1/2.
\end{equation}
Hence,
using the obvious bound $Z^{\beta,h_c(\beta),\omega}_N\ge K(N)e^{\beta\omega_N+h_c(\beta)}$, one obtains
\begin{equation}
{\ensuremath{\mathbb E}} \left[ \log Z^{\beta,h_c(\beta),\omega}_N \ | \ \sum_{x=1}^{N-1} \omega_x= v\sqrt{N-1} \right]\\
\ge
\log K(N)+h_c(\beta)+\frac{1}{2}\gep(v-u) \sqrt{N-1}.
\end{equation}
Hence integrating over $v$ one obtains (recall the value we have chosen for $u$)
\begin{multline}
{\ensuremath{\mathbb E}} \left[ \log Z^{\beta,h_c(\beta),\omega}_N\right]
\ge \log K(N)+h_c(\beta) + \frac{1}{2}\frac{1}{\sqrt{2\pi}}\int \gep(v-u)\sqrt{N-1}e^{-\frac{v^2}{2}}\dd v \\
=\log K(N)+h_c(\beta) - \frac {\gep u } {2} \sqrt{N-1} = \log K(N)+h_c(\beta)+5\sqrt{N-1}>0.
\end{multline}
This contradicts the fact that the free-energy is zero.
\end{proof}
Then we can conclude by exhibiting a finite volume bound similar to those of Lemma \ref{finitevol} for the
free energy derivative.
\begin{lemma}\label{finitevol2}
For $K$ log-convex, for any $N$ and $h$
\begin{equation}
\frac{1}{N}{\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_N\left(\sum_{n=1}^N\delta_N \right) \right]\ge \partial_h\textsc{f}(\beta,h).
\end{equation}
\end{lemma}
\begin{proof}
This is a simple consequence of the FKG inequality, as the number of contacts is an increasing function.
For $M\ge 1$ on has
\begin{multline}
{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_{MN}\left[\sum_{n=1}^{NM}\delta_{n} \right]
\ge
{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_{MN}\left[\sum_{n=1}^{NM}\delta_{n} \ | \ \delta_{iN}=1, \ \forall i\in \{1,\dots, M-1\}\right]
\\ = \sum_{i=0}^{M-1} {\ensuremath{\mathbf E}} ^{\beta,h,\theta^{iN} \omega}_{N}\left[\sum_{n=1}^{N}\delta_{n}\right],
\end{multline}
and hence taking the average
\begin{equation}
\frac{1}{NM}{\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_{MN}\left[\sum_{n=1}^{NM}\delta_{n} \right] \right]\le \frac{1}{N}{\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_{N}\left[\sum_{n=1}^{N}\delta_{n} \right]\right].
\end{equation}
The result follows by taking $M$ to infinity.
\end{proof}
\begin{proof}[Proof of Theorem \ref{mainres3}]
For a fixed $N$,
$$ h\mapsto \frac{1}{N}{\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_{N}\left[\sum_{n=1}^{N}\delta_{n} \right]\right]. $$
is a continuous function. Hence from \eqref{criticool} one can find $N$ sufficiently large and $h>h_c$ such that
\begin{equation}
\frac{1}{N}{\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h,\omega}_{N}\left[\sum_{n=1}^{N}\delta_{n} \right]\right]< c_0.
\end{equation}
By Lemma \ref{finitevol2}, this implies that $\partial_h \textsc{f} (\beta,h)<c_0$ which yields a contradiction.
Hence one must have a smooth transition.
\end{proof}
\begin{remark}
In fact the proof in this section yields a non trivial result for $\zeta<1/2$: when $K$ is $\log$-convex one has
\begin{equation}
\lim_{N\to \infty} \frac{1}{N}{\ensuremath{\mathbb E}} \left[{\ensuremath{\mathbf E}} ^{\beta,h_c(\beta),\omega}_{N}\left[\sum_{n=1}^{N}\delta_{n} \right]\right]=\lim_{h\to h_c(\beta)+} \partial_h\textsc{f}(\beta,h).
\end{equation}
In other words the contact fraction at the critical point is equal to the right-derivative of the free-energy.
\end{remark}
{\bf Acknowledgement:} The author is grateful to G. Giacomin for enlightening discussion on the subject and to N. Torri for providing access to reference \cite{cf:T14}.
|
\section{Introduction}
The goal of this article is to count the number of non-isomorphic symplectic resolutions of a symplectic quotient singularity $V / \Gamma$; where $V$ is a finite dimensional complex vector space and $\Gamma \subset \mathrm{Sp}(V)$ a finite group. A \textit{$\Q$-factorial terminalization} of $V / \Gamma$ is a projective, crepant, birational morphism
$$
\rho : Y \rightarrow V / \Gamma
$$
such that $Y$ has only $\Q$-factorial, terminal singularities. We say that $Y$ is a \textit{symplectic resolution} of $V / \Gamma$ if $Y$ is smooth. It is not always the case that the quotient admits a symplectic resolution, in fact such examples are relatively rare. However, it is a consequence of the minimal model program that $V/\Gamma$ always admits a $\Q$-factorial terminalization. Moreover, work of Namikawa shows that $V/\Gamma$ admits only finitely many $\Q$-factorial terminalizations up to isomorphism, and if one of these $\Q$-factorial terminalizations is actually smooth i.e. is a symplectic resolution, then all $\Q$-factorial terminalizations are smooth.
The main result of this paper is an explicit formula for the number of $\Q$-factorial terminalizations admitted by $V / \Gamma$. Our approach is to translate the problem into a problem about the singularities of the Calogero-Moser deformation of $V / \Gamma$. Then results about the representation theory of symplectic reflection algebras can be applied to solve the problem. Namely, the centre of the symplectic reflection algebra associated to $\Gamma$ defines a flat Poisson deformation $\CM(\Gamma) \rightarrow \mf{c}$ of $V / \Gamma$. Here the base $\mf{c}$ of the Calogero-Moser deformation is the vector space of class functions supported on the symplectic reflections in $\Gamma$. Let $\mc{Y}$ be a $\Q$-factorial terminalization of $\CM(\Gamma)$ over $\mf{c}$:
$$
\begin{tikzpicture}
\node (C1) at (-1,1) {$\mc{Y}$};
\node (C2) at (1,1) {$\CM(\Gamma)$};
\node (C3) at (0,0) {$\mf{c}$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C3);
\node [above] at (0,1) {$\boldsymbol{\rho}$};
\end{tikzpicture}
$$
The set of points $\bc$ for which the map $\boldsymbol{\rho}_{\bc} : \mc{Y}_{\bc} \rightarrow \CM_{\bc}(\Gamma)$ is an isomorphism is denoted $\mf{c}_{\reg}$, and $\mc{D} \subset \mf{c}$ the complement. In \cite{Namikawa2}, Namikawa shows that there is a finite "Weyl group" associated to any affine symplectic variety equipped with a good $\Cs$-action. In particular, we may associate to $V / \Gamma$ its Namikawa Weyl group $W$. Our main result states:
\begin{thm}\label{thm:count}
The number of pairwise non-isomorphic $\Q$-factorial terminalizations admitted by $V / \Gamma$ equals
\begin{equation}\label{eq:OSformula}
\frac{1}{|W|} \dim_{\C} H^*(\mf{c} \smallsetminus \mc{D}; \C).
\end{equation}
\end{thm}
A consequence of our results is that $\mc{D}$ is a union of hyperplanes in $\mf{c}$. This implies that $H^*(\mf{c} \smallsetminus \mc{D}; \C)$ is the Orlik-Solomon algebra associated to this hyperplane arrangement. Thus, powerful results in algebraic combinatorics can be applied to explicitly calculate the number (\ref{eq:OSformula}) in examples of interest. When $V / \Gamma$ admits a symplectic resolution, $\boldsymbol{\rho}_{\bc}$ is an isomorphism if and only if $\CM_{\bc}(\Gamma)$ is smooth i.e. $\mc{D}$ is precisely the locus of singular fibers.
There is one infinite series of groups for which it is known that the quotient $V/ \Gamma$ admits a symplectic resolution. These are the wreath product symplectic reflection groups. Let $\Gamma = \s_n \wr G$ acting on $V = \C^{2n}$; where $G$ is a finite subgroup of $SL(2,\C)$. The Weyl group associated to $G$ via the McKay correspondence is denoted $W_G$. The exponents of $W_G$ are denoted $e_1, \ds, e_{\ell}$ and $h$ denotes the Coxeter number of $W_G$.
\begin{prop}\label{prop:countWeyl}
The number of non-isomorphic symplectic resolutions of $V / \Gamma$ equals
\begin{equation}\label{eq:beautiful}
\prod_{i = 1}^{\ell} \frac{((n-1) h + e_i + 1)}{e_i + 1}
\end{equation}
\end{prop}
Formula (\ref{eq:beautiful}) plays an important role in the theory of generalized Catalan combinatorics associated to Weyl groups.
In addition to the above infinite series, it is known that there are two exceptional groups that admit symplectic resolutions. These are $Q_8 \times_{\Z_2} D_8$ and $G_4$, both acting on a four-dimensional symplectic vector space; it seems likely that these make up all groups admitting symplectic resolutions \cite{SchUn}. In the case of $G_4$, Lehn and Sorger explicitly constructed a pair of non-isomorphic symplectic resolutions of $V / \Gamma$. Our results show that these are the only symplectic resolutions of this quotient. In the case of $Q_8 \times_{\Z_2} D_8$, a computer calculation shows that $\dim_{\C} H^*(\mf{c} \smallsetminus \mc{D}; \C) = 2592$, implying that the quotient singularity admits $81$ distinct symplectic resolutions. Recently, these $81$ symplectic resolutions have been explicitly constructed by M. Donten-Bury and J. A. Wi\'sniewski \cite{Wisniewski}. They also show that these $81$ resolutions are all possible resolution up to isomorphism.
\subsection{Universal vs. Calogero-Moser deformations}\label{sec:uvsp}
The key to proving Theorem \ref{thm:count} is to make a precise comparison between the formally universal Poisson deformation $\mc{X}$ of $V / \Gamma$ and the Calogero-Moser deformation $\CM(\Gamma)$. As noted above, the base of the Calogero-Moser space is the space $\mf{c}$ of class functions on $\Gamma$ supported on the subset of symplectic reflections. On the other hand, Namikawa has shown that the base of the universal deformation $\mc{X}$ is $H^2(Y;\C) / W$. Thus, there exists a morphism $\mf{c} \rightarrow H^2(Y;\C) / W$ such that
$$
\CM(\Gamma) \simeq \mc{X} \times_{H^2(Y;\C) / W} \mf{c}.
$$
Our main result, Theorem \ref{thm:main}, is an explicit description of the morphism $\mf{c} \rightarrow H^2(Y;\C) / W$. In order to precisely state our results, we introduce some additional notation.
A subgroup $\uGamma$ of $\Gamma$ is \textit{parabolic} if it is the stabilizer of some vector $v \in V$. The rank of $\uGamma$ is defined to be $\frac{1}{2} \left(\dim V - \dim V^{\uGamma} \right)$, and we say that $\uGamma$ is \textit{minimal} if it has rank one. In this case $\uGamma$ is isomorphic to a finite subgroup of $SL(2,\C)$. The set of $\Gamma$-conjugacy classes of minimal parabolic subgroups is denoted $\mc{B}$. The variety $V / \Gamma$ is stratified by finitely many symplectic leaves and those leaves $\mc{L}$ whose dimension is $\dim V - 2$ are naturally labeled by the elements of $\mc{B}$. For each $B \in \mc{B}$, we fix a representative $\uGamma$ in $B$ and write $\widetilde{\Xi}(B)$ for the normalizer of $\uGamma$ in $\Gamma$. The quotient $\widetilde{\Xi}(B) / \uGamma$ is denoted $\Xi(B)$. Via the McKay correspondence, there is associated to $\uGamma \subset SL(2,\C)$ a Weyl group $(W(B),\h_B)$, of simply laced type. As explained in section \ref{sec:McKay}, there is a natural linear action of $\Xi(B)$ on $\h_B$. We fix $\mf{a}_B := (\h^*_B)^{\Xi(B)}$. The centralizer $W_B$ of $\Xi(B)$ in $W(B)$ acts on $\mf{a}_B$. We fix a $\Q$-factorial terminalization $\rho : Y \rightarrow V / \Gamma$ of $V / \Gamma$.
\begin{thm}\label{prop:Weylgroup}
The Namikawa Weyl group associated to $V/\Gamma$ is $W := \prod_{B \in \mc{B}} W_B$ acting on
$$
H^2(Y;\C) \simeq \prod_{B \in \mc{B}} \mf{a}_B.
$$
\end{thm}
As noted above, the Calogero-Moser deformation plays a key role in our results. Associated to the pair $(V,\Gamma)$ is the \textit{symplectic reflection algebra} $\mathbf{H}(\Gamma)$ at $t = 0$, as introduced by Etingof and Ginzburg \cite{EG}. This is a non-commutative $\C[\mf{c}]$-algebra, free over $\C[\mf{c}]$, such that the quotient $\mathbf{H}(\Gamma) / \langle \C[\mf{c}]_+ \rangle$ is isomorphic to the skew-group algebra $\C[V] \rtimes \Gamma$. Let $e$ denote the trivial idempotent in $\C \Gamma$, so that $e(\C[V] \rtimes \Gamma)e \simeq \C[V]^{\Gamma}$. The algebra $e \mathbf{H}(\Gamma) e$ is a commutative Poisson algebra, again free over $\C[\mf{c}]$, such that
$$
e \mathbf{H}(\Gamma) e / \langle \C[\mf{c}]_+ \rangle \simeq \C[V]^{\Gamma},
$$
as Poisson algebras. Thus, $\vartheta : \CM(\Gamma) := \mathop{\rm Spec}\nolimits e \mathbf{H} e \rightarrow \mf{c}$ is a flat Poisson deformation of $V / \Gamma$. We call $\CM(\Gamma)$ the \textit{Calogero-Moser deformation} of $V / \Gamma$. The key result at the heart of this paper is the following theorem, which makes explicit the relation between the deformations $\mc{X}$ and $\CM(\Gamma)$ of $V / \Gamma$.
\begin{thm}\label{thm:main}
The McKay correspondence defines a $W$-equivariant isomorphism $\mf{c} \simeq H^2(Y;\C)$ such that the Calogero-Moser deformation $\CM(\Gamma) \rightarrow \mf{c}$ is isomorphic to the pull-back along the quotient map $H^2(Y;\C) \rightarrow H^2(Y;\C) / W$ of the formally universal Poisson deformation $\mc{X} \rightarrow H^2(Y;\C) / W$.
\end{thm}
In particular, Theorem \ref{thm:main} implies that Conjecture 1.9 of \cite{GK} is true. Results of Namikawa \cite{Namikawa3} on the birational geometry of $Y$ show that the number of $\Q$-factorial terminalizations of $V / \Gamma$ can be computed by counting the number of connected components in the complement to a (finite) real hyperplane arrangement in $H^2(Y;\R)$. Theorem \ref{thm:main} allows us to identify the complexification of this hyperplane arrangement with the set $\mc{D}$. Then Theorem \ref{thm:count} can be deduced from Theorem \ref{thm:main} using standard results from the theory of hyperplane arrangements. We note an immediate corollary of Theorem \ref{thm:main}.
\begin{cor}
Let $\Gamma_0$ be the normal subgroup of $\Gamma$ generated by all symplectic reflections. Then the number of $\Q$-factorial terminalizations of $V / \Gamma$ equals the number of $\Q$-factorial terminalizations admitted by $V / \Gamma_0$.
\end{cor}
Thus, if $\Gamma_0 = \{ 1 \}$, then $V / \Gamma$ is the unique $\Q$-factorial terminalization of $V / \Gamma$. Let $\mc{Y}$ be the formally universal Poisson deformation of the terminialization $Y$. Then $\mc{Y}$ is projective over its affinization $\mc{Y}^{\mathrm{aff}} := \mathop{\rm Spec}\nolimits \Gamma(\mc{Y},\mc{O})$.
\begin{cor}\label{cor:affinization}
Then there exists an isomorphism of Poisson $H^2(Y;\C)$-schemes $\mc{Y}^{\mathrm{aff}} \simeq \CM(\Gamma)$.
\end{cor}
The birational geometry of $\Q$-factorial terminalizations of $V / \Gamma$ can also be used to deduce results about the Calogero-Moser deformation. Namely, the following is a partial answer to Question 9.8.4 by Bonnaf\'e and Rouquier \cite{BonnafeRouquier}.
\begin{cor}\label{cor:singCal}
Let $\mc{D}' \subset \mf{c}$ be the locus over which the fibers of the Calogero-Moser deformation $\CM(\Gamma)$ are singular. Then $\mc{D}'$ is either a finite union of hyperplanes, or the whole of $\mf{c}$.
\end{cor}
\subsection{Outline of paper}
In section \ref{sec:proof2}, we give a proof of Theorem \ref{prop:Weylgroup}. Section \ref{sec:CMdef} is devoted to the proof of Theorem \ref{thm:main}. The proof of Corollary \ref{cor:singCal} is given in section \ref{sec:BR}. Then, our main result Theorem \ref{thm:count} is proven in section \ref{sec:proof1}. Finally, we consider specific examples in sections \ref{sec:wreath} and \ref{sec:except}, where formula (\ref{eq:beautiful}) of Proposition \ref{prop:countWeyl} is derived.
\begin{remark}
Throughout, the cohomology group $H^i(Y;\C)$ stands for the singular cohomology of underlying reduced variety, equipped with the \textit{analytic} topology.
\end{remark}
\section{Namikawa's Weyl group}\label{sec:proof2}
In this section, we describe Namikawa's Weyl group associated to $V / \Gamma$, thus confirming Theorem \ref{prop:Weylgroup}.
\subsection{}\label{sec:McKay} The symplectic leaves in $V / \Gamma$ are labeled by $\Gamma$-conjugacy classes of parabolic subgroups of $\Gamma$. Let $\underline{\Gamma}$ be a parabolic subgroup. Then the leaf $\mc{L}$ labeled by $\underline{\Gamma}$ is the image under $\pi : V \rightarrow V / \Gamma$ of the set $\{ v \in V \ |\ \Gamma_v = \underline{\Gamma} \}$. If $( V / \Gamma)_{\le 1}$ is the open subset consisting of the open symplectic leaf and all leaves $\mc{L}$ of dimension $\dim V - 2$, then we write $\Yle := \rho^{-1}(( V / \Gamma)_{\le 1})$. The open subset $\Yle$ is contained in the smooth locus of $Y$.
As in section \ref{sec:uvsp}, we fix $B \in \mc{B}$, $\uGamma$ the corresponding minimal parabolic in $\Gamma$ etc. Let $V_0$ denote the complementary $\uGamma$-module to $V^{\uGamma}$ in $V$; $V_0$ is a two-dimensional symplectic subspace. The open subset of $V^{\uGamma}$ consisting of all points whose stabilizer under $\Gamma$ equals $\uGamma$ is denoted $U$. The group $\Xi(B)$ acts freely on $U \times V_0 / \uGamma$ and the quotient map $\pi$ induces a Galois covering $\sigma : U \times V_0 / \uGamma \rightarrow V / \Gamma$ onto its image, with Galois group $\Xi(B)$.
We choose $b \in U$ and set $p = \pi(b)$. Then $\pi(\{ b \} \times V_0) \simeq V_0 / \uGamma$ is a closed subvariety of $V / \Gamma$. Let $Y_B = \rho^{-1}(V_0 / \uGamma)$ in $Y$, so that $Y_B \subset \Yle$ and $\rho : Y_B \rightarrow V_0 / \uGamma$ is a minimal resolution of singularities. Let $F$ be the exceptional locus of this minimal resolution and $\Irr (F)$ the set of exceptional divisors. Recall that $W(B)$ is the Weyl group associated to $\uGamma$. Let $\Delta_B \subset \h_B^*$ be a set of simple roots. The set of isomorphism classes of \textit{non-trivial} irreducible $\uGamma$-modules is denoted $\Irr (\uGamma)$. By \cite[Theorem 2.2 (i)]{GonzalezSprinbergVerdier}, the McKay correspondence is the pair of bijections
\begin{equation}\label{eq:mckaybij}
\Delta_B {\;\stackrel{_\sim}{\to}\;} \Irr (\uGamma) {\;\stackrel{_\sim}{\to}\;} \Irr (F), \quad \alpha \mapsto \rho(\alpha) \mapsto D_{\rho(\alpha)},
\end{equation}
uniquely defined by the condition
\begin{equation}\label{eq:pairings}
(D_{\rho(\alpha)},D_{\rho(\beta)}) = \dim \Hom_{\uGamma}( V_0 \o \rho(\alpha), \rho(\beta)) = - \langle \alpha, \beta \rangle,
\end{equation}
where $( - , - )$ is the intersection pairing and $\langle - , - \rangle$ the Killing form. There is a natural representation theoretic action of $\Xi(B)$ on the set $\Irr (\uGamma) $. For $\lambda \in \Irr (\uGamma)$ and $x \in \Xi(B)$, we have $x \cdot \lambda = {}^x \lambda$, where ${}^x \lambda$ is the $\uGamma$-module, which as a vector space equals $\lambda$, with action $g \cdot v = xgx^{-1} v$ for all $v \in \lambda$. The identity (\ref{eq:pairings}) implies that the induced action of $\Xi(B)$ on $\Delta_B$ is via Dynkin diagram automorphism.
\subsection{} The group $\Xi(B)$ also acts naturally on $H^2(Y_B; \C)$ as follows. Since the decomposition $V = V^{\uGamma} \oplus V_0$ is as $\widetilde{\Xi}(B)$-modules, $\Xi(B)$ acts on $V_0 / \uGamma \subset V / \uGamma$. There is a unique lift of this action to the resolution $Y_B$, as can be seen from the explicit construction of $Y_B$ as $\mathrm{Hilb}^{\uGamma}(\C^2)$, the dominant component of $\uGamma$-$\mathrm{Hilb}(\C^2)$; see \cite{CBKleinian}. Thus, there is an induced action of $\Xi(B)$ on $H^2(Y_B;\C)$.
Recall that each divisor $D \in \Irr (F)$ is a rational curve with self-intersection $-2$. For $D \in \Irr (F)$, let $L_D$ denote the corresponding line bundle on $Y_B$ such that $L_D |_D \simeq \mc{O}_D(-1)$ and $L_D |_{D'} = \mc{O}_{D'}$ for $D' \neq D$. The following is a well-known part of the McKay correspondence, but we sketch a proof since we were unable to find a suitable reference.
\begin{lem}\label{lem:twodimHtwo}
For $\mc{L} \in \Pic(Y_B)$, let $c_1(\mc{L})$ denote its Chern character in $H^2(Y_B;\C)$.
\begin{enumerate}
\item The Chern characters $c_1(L_D)$, for $D \in \Irr (F)$, are a basis of $H^2(Y_B;\C)$.
\item The induced isomorphism
$$
\h^*_B {\;\stackrel{_\sim}{\to}\;} H^2(Y_B;\C), \quad \alpha \mapsto c_1 \left( L_{D_{\rho(\alpha)}} \right)
$$
is $\Xi(B)$-equivariant.
\end{enumerate}
\end{lem}
\begin{proof}
Both statements will be proven simultaneously. We have defined the action of $\Xi(B)$ on $\Delta_B$ such that the bijection $\Delta_B {\;\stackrel{_\sim}{\to}\;} \Irr (\uGamma)$ is equivariant. Since the action of $\Xi(B)$ on $V_0 / \uGamma$ fixes the singular point, $\Xi(B)$ acts on $F$, permuting its irreducible components. Thus, there is a geometric action of $\Xi(B)$ on $\Irr (F)$. It follows from the beautiful interpretation of the bijection $\Irr (\uGamma) {\;\stackrel{_\sim}{\to}\;} \Irr (F)$ given in \cite{CBKleinian} that this bijection is $\Xi(B)$-equivariant; see also \cite[Section 6.2]{SchUn}. Thus, it suffices to check that the Chern characters $c_1(L_D)$, for $D \in \Irr (F)$, are a basis of $H^2(Y_B;\C)$ such that $x \cdot c_1(L_D) = c_1(L_{x \cdot D})$ for all $x \in \Xi(B)$.
The action of $\Xi(B)$ on $Y_B$ commutes with the natural conic $\Cs$-action. Therefore, \cite[Proposition 4.3.1]{SlodowyFourLectures} shows that the embedding $F \hookrightarrow Y_B$ induces, by restriction, a $\Xi(B)$-equivariant isomorphism $H^2(Y_B,\C) {\;\stackrel{_\sim}{\to}\;} H^2(F,\C)$. Now, by the Mayer-Viratoris long exact sequence, the embeddings $D \hookrightarrow F$ identify $H^2(F;\C)$ with $\bigoplus_{D \in \Irr (F)} H^2(D;\C)$. Under the identification
$$
H^2(Y_B,\C) {\;\stackrel{_\sim}{\to}\;} \bigoplus_{D \in \Irr (F)} H^2(D;\C)
$$
the group $\Xi(B)$ acts by permuting the (one-dimensional) summands of the right-hand side. On the other hand, the image of $c_1(L_D)$ in $H^2(D';\C)$ is either a basis element if $D = D'$, or zero if $D \neq D'$, since $c_1 \left( \mc{O}_{\mathbb{P}^1} \right) = 0$ in $H^2(\mathbb{P}^1;\C)$. The claims of the lemma follow.
\end{proof}
Define $Z$ to be the fiber product
$$
\begin{tikzpicture}
\node (C1) at (-1.5,1.5) {$Z$};
\node (C2) at (1.5,1.5) {$Y$};
\node (C3) at (-1.5,0) {$U \times V_0 / \uGamma$};
\node (C4) at (1.5,0) {$V / \Gamma$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C3) -- (C4);
\node [above] at (0,1.5) {$\sigma'$};
\node [above] at (0,0) {$\sigma$};
\end{tikzpicture}
$$
Since $\sigma$ is \'etale, $\sigma'$ is also \'etale by base change. The following is based on \cite[Proposition 5.2]{KaledinDynkinArxiv}.
\begin{prop}\label{prop:Ziso}
There is a $\Xi(B)$-equivariant isomorphism $U \times Y_B {\;\stackrel{_\sim}{\to}\;} Z$.
\end{prop}
\begin{proof}
Set $U_0 = U \times V_0 / \uGamma$. Since $\sigma$ is \'etale, and $U_0 \times_{V/\Gamma} Y = U_0 \times_{V/\Gamma} \Yle$, the fiber product $Z$ is a smooth variety. Projective base change implies that it is projective over $U_0$. If $(V_0 / \uGamma)_{\reg}$ is the smooth locus of $V_0 / \uGamma$ then $U \times (V_0 / \uGamma)_{\reg}$ is an open subset of $V^{\uGamma} \times (V_0 / \uGamma)_{\reg}$ with compliment of codimension at least two. Hence $\Pic(U \times (V_0 / \uGamma)_{\reg}) \simeq \Pic((V_0 / \uGamma)_{\reg})$ is torsion. Therefore, the proof of \cite[Lemma 5.1]{KaledinDynkinArxiv} implies that there is a sheaf of ideals $\mc{E} \subset \mc{O}_{U_0}$ and an isomorphism
$$
Z {\;\stackrel{_\sim}{\to}\;} \mathrm{Bl}(U_0, \mc{E}),
$$
of varieties projective over $U_0$, where $\mathrm{Bl}(U_0, \mc{E})$ is the blowup of $U_0$ along $\mc{E}$. Since the line bundle on $Z$, ample relative to $U_0$, used to embed $Z$ in $\mathbb{P}^N_{U_0}$ is the pullback of a line bundle on $Y$, ample relative to $V / \Gamma$, the identification $Z \simeq \mathrm{Bl}(U_0, \mc{E})$ is $\Xi(B)$-equivariant i.e. $\mc{E}$ is $\Xi(B)$-stable. To show that $ \mathrm{Bl}(U_0, \mc{E}) \simeq U \times Y_B$, we follow the proof of \cite[Proposition 5.2]{KaledinDynkinArxiv}. Based on the argument given there, it is clear that it suffices to show that all the vector fields $\mf{t}_v$ on $U_0$ coming from the constant coefficient vector fields $v \in V^{\uGamma}$ admit lifts to $Z$.
The projective morphism $\rho : \Yle \rightarrow \rho(\Yle)$ is semi-small since $\Yle$ is a symplectic manifold \cite[Theorem 3.2]{FuSurvey}. Therefore, since the map $\sigma : U_0 \rightarrow V / \Gamma$ is finite onto its image, the map $Z \rightarrow U_0$ is also semi-small. Moreover, by \'etale base change, the fact that the canonical bundle on $\Yle$ is trivial implies that the canonical bundle on $Z$ is trivial too. Therefore, the required lifting follows from \cite[Lemma 5.3]{GK}.
\end{proof}
\begin{lem}\label{lem:fundamental}
The fundamental group $\pi_1(\mc{L})$ of $\mc{L}$ equals $\Xi$.
\end{lem}
\begin{proof}
The leaf $\mc{L}$ is the image under $\sigma$ of $U \times \{ 0 \} \subset U \times V_0 / \uGamma$. Thus, $\mc{L} \simeq U / \Xi$. Since $\Xi$ acts freely on $U$ this implies that we have a short exact sequence $1 \rightarrow \pi_1(U) \rightarrow \pi_1(\mc{L}) \rightarrow \Xi \rightarrow 1$. Hence, it suffices to show that $\pi_1(U)$ is trivial. The complement of $U$ in $V^{\uGamma}$ is the union of subspaces $V^{\uGamma} \cap V^{\uGamma'}$, where $\uGamma'$ is a parabolic subgroup of $\Gamma$ such that $V^{\uGamma} \cap V^{\uGamma'}$ is a proper subspace of $ V^{\uGamma}$. We may assume that $\uGamma \subsetneq \uGamma'$ so that $V^{\uGamma'} \subsetneq V^{\uGamma}$. But $V^{\uGamma'} $ is a symplectic subspace of $V$. Thus, $\dim V^{\uGamma'} < \dim V^{\uGamma} - 1$. Hence, the compliment of $U$ in $V^{\uGamma}$ has codimension at least two, implying that $\pi_1(U)$ is trivial.
\end{proof}
If $(V / \Gamma)_0$ is the open leaf in $V / \Gamma$, then we denote by $Y_0$ the preimage of $(V / \Gamma)_0$ under $\rho$. The map $\rho$ is an isomorphism over $Y_0$.
\begin{lem}\label{lem:Hvanish}
For $0 < i < 4$, the cohomology groups $H^i(U;\C)$ and $H^i(Y_0;\C)$ are zero.
\end{lem}
\begin{proof}
As shown in the proof of Lemma \ref{lem:fundamental}, the compliment $C$ to $U$ in $V^{\uGamma}$ has complex codimension at least two. Therefore,
$$
H^{\mathrm{BM}}_j(C;\C) = H^{2 \dim_{\C} V^{\uGamma} - i}\left( V^{\uGamma}_{\R},V^{\uGamma}_{\R} \smallsetminus C; \C \right) = 0, \quad \forall \ j > 2 \dim_{\C} C,
$$
where $\mathrm{BM}$ indicates Borel-Moore homology. This implies that $H^{i}\left(V^{\uGamma}_{\R},V^{\uGamma}_{\R} \smallsetminus C; \C\right) = 0$ for $i < 4$. Since $H^i\left(V^{\uGamma}_{\R};\C\right) = 0$ for $i > 0$, the first claim follows from the long exact sequence in relative cohomology. For the second claim, we note first that if $V_{\reg}$ is the open subset of $V$ on which $\Gamma$ acts freely, then $\rho$ restricts to an isomorphism $Y_0 {\;\stackrel{_\sim}{\to}\;} \pi(V_{\reg})$. On $V_{\reg}$, the map $\pi$ is a covering with Galois group $\Gamma$. Therefore, by \cite[Proposition 3.G.1]{Hatcher}, it suffices to show that $H^i(V_{\reg};\C) = 0$ for $0 < i < 4$. Again, this follows from the fact that the compliment to $V_{\reg}$ in $V$ has complex codimension at least $2$.
\end{proof}
\begin{lem}\label{lem:H2facts}
Fix a $\Q$-factorial terminalization $\rho : Y \rightarrow V/\Gamma$. Then $H^2(Y_{\le 1};\R) = H^2(Y;\R)$.
\end{lem}
\begin{proof}
Let $Y^o$ denote the smooth locus of $Y$ and $X^o$ its image under $\rho$. Since $Y$ has terminal singularities, the compliment of $Y^o$ in $Y$ has codimension at least $4$ \cite{NamikawaNote}. Then \cite[Theorem 3.2]{FuSurvey} says that $\rho$ restricted to $Y^o$ is a semi-small map. Therefore $Y \smallsetminus \Yle$ has codimension at least $4$ in $Y$. The lemma follows from the argument given in the proof of Lemma \ref{lem:Hvanish} above.
\end{proof}
\begin{prop}\label{prop:H2iso}
The restriction maps $H^2(Y; \C) \rightarrow H^2(Y_B;\C)$ induce an isomorphism
$$
H^2(Y; \C) {\;\stackrel{_\sim}{\to}\;} \bigoplus_{B \in \mc{B}} H^2(Y_B;\C)^{\Xi(B)}.
$$
\end{prop}
\begin{proof}
By Lemma \ref{lem:H2facts}, it suffices to show that the restriction maps $H^2(\Yle; \C) \rightarrow H^2(Y_B;\C)$ induce an isomorphism
$$
H^2(\Yle; \C) {\;\stackrel{_\sim}{\to}\;} \bigoplus_{B \in \mc{B}} H^2(Y_B;\C)^{\Xi(B)}.
$$
Let $Y(B) \subset Y$ be the open set $\rho^{-1}(\sigma(U \times V_0 / \uGamma))$. Then for $B \neq B'$ in $\mc{B}$, we have $Y(B) \cap Y(B') = Y_0$ and $\Yle = \bigcup_{B} Y(B)$. We claim that restriction defines an isomorphism
$$
H^2(\Yle;\C) {\;\stackrel{_\sim}{\to}\;} \bigoplus_{B \in \mc{B}} H^2(Y(B);\C).
$$
This follows from the Mayer-Vietoris sequence by induction on $|\mc{B}|$, using the fact that $H^i(Y_0; \C) = 0$ for $0 < i < 4$ by Lemma \ref{lem:Hvanish}. Therefore we are reduced to showing that restriction $ H^2(Y(B);\C) \rightarrow H^2(Y_B;\C)$ is injective with image $H^2(Y_B;\C)^{\Xi(B)}$.
Recall that we identified $Y_B$ with a closed subset of $Y$ by first fixing $b \in U$ and identifying $V_0 / \uGamma$ with $\sigma(\{ b \} \times V_0 / \uGamma)$ in $V / \Gamma$. Therefore, the closed embedding $Y_B \hookrightarrow Y$ factors as $Y_B \stackrel{j}{\rightarrow} Z \stackrel{\sigma}{\rightarrow} Y$, where $j$ is the closed embedding $u \mapsto (b,u)$ in $Z \simeq U \times Y_B$ of Proposition \ref{prop:Ziso}. Then Lemma \ref{lem:Hvanish} and the Kunneth formula imply that $j$ induces an identification $H^2(Z;\C) = H^2(Y_B;\C)$.
The image of $Z$ in $Y$ under the natural map $\sigma' : Z \rightarrow Y$ equals $Y(B)$. Recall that this map is just the quotient map for the free action of $\Xi(B)$ on $Z$. Therefore, by \cite[Proposition 3.G.1]{Hatcher}, pullback along $\sigma'$ is injective with image $H^2(Z;\C)^{\Xi(B)} = H^2(Y_B;\C)^{\Xi(B)}$.
\end{proof}
Theorem \ref{prop:Weylgroup} is now a direct consequence of Proposition \ref{prop:H2iso} and the proof of Theorem 1.1 of \cite{Namikawa2} given in \textit{loc. cit.} In particular, the identification $W \simeq \prod_{B \in \mc{B}} W_B$ follows from Lemma 1.2 of \textit{loc. cit.}
\begin{remark}
Let $\mc{L}$ be the leaf in $V / \Gamma$ labeled by $B \in \mc{B}$. The restriction of $\rho$ to $\rho^{-1}(\mc{L})$ is (in the analytic topology) a fiber bundle with fiber $F$. Therefore, by Lemma \ref{lem:fundamental}, the action of $\Xi(B)$ on $H^2(Y_B; \C) \simeq H^2(F;\C)$ is the monodromy action of $\pi_1(\mc{L}) = \Xi(B)$.
\end{remark}
\section{Calogero-Moser deformations}\label{sec:CMdef}
Our approach to the proof of Theorem \ref{thm:main} will be by analogy with the proof of \cite[Theorem 1.1]{Namikawa2}. As in previous sections we fix a $\Q$-factorial terminalization $\rho : Y \rightarrow V / \Gamma$.
\subsection{Formally universal Poisson deformations} Recall from Lemma \ref{lem:H2facts} that the cohomology group $H^2(\Yle;\C)$ equals $H^2(Y;\C)$. By \cite[Theorem 5.5]{Namikawa} and \cite[Theorem 1.1]{Namikawa}, there are flat Poisson deformations $\nu : \mc{X} \rightarrow H^2(Y;\C) / W$, and $\boldsymbol{\nu} : \mc{Y} \rightarrow H^2(Y;\C)$, of $V / \Gamma$ and $Y$ respectively, such that the diagram
\begin{equation}\label{eq:commdiNam}
\begin{tikzpicture}
\node (C1) at (-1.5,1.5) {$\mc{Y}$};
\node (C2) at (1.5,1.5) {$\mc{X}$};
\node (C3) at (-1.5,0) {$H^2(Y;\C)$};
\node (C4) at (1.5,0) {$H^2(Y;\C) / W$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C3) -- (C4);
\node [left] at (-1.5,0.75) {$\boldsymbol{\nu}$};
\node [right] at (1.5,0.75) {$\nu$};
\end{tikzpicture}
\end{equation}
is commutative. Moreover, the natural conic action of the torus $\Cs$ on $V/\Gamma$ lifts to the flat families $\mc{X} \rightarrow H^2(Y;\C) / W$ and $\mc{Y} \rightarrow H^2(Y;\C)$ in such a way that $\lambda \cdot h = \lambda^2 h$ for all $h \in H^2(Y;\C)^* \subset \C[H^2(Y;\C)]$ and $\lambda \in \Cs$. The maps in (\ref{eq:commdiNam}) are equivariant for this action.
The flat Poisson deformation $\mc{X} \rightarrow H^2(Y;\C) / W$ is universal in the following sense. If $\mc{X}' \rightarrow T$ is a flat Poisson deformation of $V/\Gamma$ over a local Artinian $\C$-scheme, then there exists a unique morphism $T \rightarrow H^2(Y;\C) / W$ such that $\mc{X}' \simeq \mc{X} \times_{H^2(Y;\C) / W} T$ as Poisson schemes over $T$. Notice that the map $T \rightarrow H^2(Y;\C) / W$ necessarily factors through the completion of $H^2(Y;\C) / W$ at zero. Thus, $\mc{X} \rightarrow H^2(Y;\C) / W$ is said to be the \textit{formally universal} Poisson deformation of $V/\Gamma$. Similarly, the deformation $\mc{Y} \rightarrow H^2(Y;\C)$ of $Y$ is formally universal.
As written, formally universal Poisson deformations of $V/\Gamma$ are clearly not unique, since the definition only involves the completion of the base of the deformation at the special fiber. However, the torus $\Cs$ acts naturally on the ring of functions on this formal neighborhood of the special fiber, see \cite[Section 5.4]{Namikawa}, and $H^2(Y;\C) / W$ is unique in the sense that it is the globalization, as explained in section \ref{sec:global} below, of the formal neighborhood.
\subsection{Symplectic reflection algebras}\label{sec:SRA} The set of \textit{symplectic reflections} $\mc{S}$ in $\Gamma$ is the set of all elements $s$ such that $\mathrm{rk}_V (1 - s) = 2$. Let $S_1, \ds, S_r$ be the $\Gamma$-conjugacy classes in $\mc{S}$ and $\bc_1,\ds ,\bc_r$ the characteristic functions on $\mc{S}$ such that $\bc_i(s) = 1$ if $s \in S_i$, and is zero otherwise. The linear span of $\bc_1,\ds ,\bc_r$ is denoted $\mf{c}$. Since we do not require the explicit definition of the symplectic reflection algebras $\mathbf{H}(\Gamma)$, and will only use results from \cite{LosevSRAComplete} about them, we refer the read to \textit{loc. cit.} for their definition. Recall that $\uGamma_B$ is a representative in the conjugacy class $B$ of minimal parabolic subgroups of $\Gamma$.
\begin{lem}\label{lem:bij1}
The natural map $\zeta : \bigsqcup_{B \in \mc{B}} (\uGamma_B \smallsetminus \{ 1 \}) / \widetilde{\Xi}(B) \longrightarrow \mc{S} / \Gamma$ is a bijection.
\end{lem}
Let $\mf{c}_B$ be the subspace of $\mf{c}$ spanned by all $\bc_i$ such that $S_i \cap \uGamma_B \neq \emptyset$. Lemma \ref{lem:bij1} implies that $\mf{c} = \bigoplus_{B \in \mc{B}} \mf{c}_B$. Choose $B \in \mc{B}$. Via the McKay correspondence (\ref{eq:mckaybij}), an element $h \in \h^*_B$ can be considered as a linear combination of the non-trivial characters of $\uGamma_B$. In other words, it is a class function on $\uGamma_B$. Hence an element of $\mf{a}_B = (\h_B^*)^{\Xi(B)}$ is a $\widetilde{\Xi}(B)$-equivariant function $\uGamma_B \rightarrow \C$, where $\widetilde{\Xi}(B)$ acts trivially on $\C$. Thus, we may define an isomorphism
\begin{equation}\label{eq:mcKay}
\varpi : \mf{c} = \bigoplus_{B \in \mc{B}} \mf{c}_B \stackrel{\sim}{\longrightarrow} \bigoplus_{B \in \mc{B}} \mf{a}_B, \quad \mf{c}_B \ni \bc \ \mapsto \ (g \mapsto \bc(\zeta(g))).
\end{equation}
As explained in section \ref{sec:uvsp}, the spherical subalgebra $e \mathbf{H}(\Gamma) e$ is a commutative $\C[\mf{c}]$-subalgebra, equipped with a natural Poisson structure, such that the flat family $\vartheta : \CM(\Gamma) \rightarrow \mf{c}$ is a Poisson deformation of $V / \Gamma$.
\subsection{Globalization}\label{sec:global}
Suppose we have two conic affine varieties $X$ and $Y$ i.e. $\C[X]$ and $\C[Y]$ are positively graded algebras with degree zero part equal to $\C$, and an equivariant morphism $\gamma : X \rightarrow Y$. Let $X^{\wedge}$ and $Y^{\wedge}$ denote the completions of $X$ and $Y$ respectively at the $\Cs$-fixed point. As shown in \cite[Lemma (A.2)]{NamikawaPoissondeformations}, $\C[X]$ is the ring of $\Cs$-locally finite ($=$ rational) vectors in $\C[X^{\wedge}]$. We say that $\gamma$ is the \textit{globalization} of $\hat{\gamma} : X^{\wedge} \rightarrow Y^{\wedge}$ if, under the identification of $\C[X]$ with rational vectors in $\C[X^{\wedge}]$ and similarly for $\C[Y] \subset \C[Y^{\wedge}]$, $\gamma$ is just the restriction of $\hat{\gamma}$; see \cite[Appendix]{NamikawaPoissondeformations}.
If $\mc{X}^{\wedge}$ is the completion of $\mc{X}$ along the closed subvariety $V / \Gamma$ and $(H^2(Y;\C) / W)^{\wedge}$ the completion of $H^2(Y;\C) / W$ at $o$, then $\nu$ is the globalization of the induced map of formal schemes $\mc{X}^{\wedge} \rightarrow (H^2(Y;\C) / W)^{\wedge}$. The latter is the universal Poisson deformation of $V / \Gamma$ in the category of pro-Artinian local $\C$-algebras. The analogous statement holds for $\mc{Y} \rightarrow H^2(Y;\C)$; see \cite[Section 5]{Namikawa}.
The Calogero-Moser deformation $\vartheta : \CM(\Gamma) \rightarrow \mf{c}$ of $V / \Gamma$ is also $\Cs$-equivariant, where $\Cs$ acts on $\mf{c}^* \subset \C[\mf{c}]$ by $\lambda \cdot \bc = \lambda^2 \bc$. This is a consequence of the fact that the symplectic reflection algebra $\mathbf{H}(\Gamma)$ is naturally $\N$-graded, such that $\mf{c}^* \subset \mathbf{H}(\Gamma)$ has degree two, $V^*$ has degree one and $\Gamma$ sits in degree zero. Moreover, if $\mathbf{H}(\Gamma)^{\wedge}$ is the completion of $\mathbf{H}(\Gamma)$ along the two-sided ideal generated by $\mf{c}^*$, then one can identify $\mathbf{H}(\Gamma)$ with the subalgebra of $\mathbf{H}(\Gamma)^{\wedge}$ of rational vectors. This implies that $\CM(\Gamma) \rightarrow \mf{c}$ is the globalization of $\CM(\Gamma)^{\wedge} \rightarrow \widehat{\mf{c}}$, where $\CM(\Gamma)^{\wedge}$ is the completion of $\CM(\Gamma)$ along $V / \Gamma$ and $\widehat{\mf{c}}$ the completion of $\mf{c}$ at zero. Hence, there exists a unique $\Cs$-equivariant morphism $\hat{\alpha} : \widehat{\mf{c}} \rightarrow (H^2(Y;\C) / W)^{\wedge}$ such that $\CM(\Gamma)^{\wedge} \simeq \widehat{\mf{c}} \times_{(H^2(Y;\C) / W)^{\wedge}} \mc{X}^{\wedge}$. This implies:
\begin{lem}\label{lem:classmap}
There exists a unique $\Cs$-equivariant map $\alpha : \mf{c} \rightarrow H^2(Y;\C) / W$ such that
$$
\CM(\Gamma) \simeq \mf{c} \times_{H^2(Y;\C) / W} \mc{X}.
$$
\end{lem}
On the other hand, the linear isomorphism (\ref{eq:mcKay}) together with the quotient map $H^2(Y;\C) \rightarrow H^2(Y;\C) / W$ defines a map $\beta : \mf{c} \rightarrow H^2(Y;\C) / W$ which is clearly also the globalization of $\hat{\beta} : \widehat{\mf{c}} \rightarrow (H^2(Y;\C) / W)^{\wedge}$. Theorem \ref{thm:main} is claiming that $\alpha = \beta$. It suffices instead to show that
$$
\hat{\alpha} = \hat{\beta} : \widehat{\mf{c}} \rightarrow (H^2(Y;\C) / W)^{\wedge}.
$$
This will be our goal for the remainder of the section.
\subsection{Kleinian singularities} In this section we consider the case $\dim V = 2$, and hence $\Gamma$ is a Kleinian group. As noted in section \ref{sec:uvsp}, associated to $\Gamma$ via the McKay correspondence is a Weyl group $(W,\h)$. Let $Y$ be the minimal resolution of $V / \Gamma$. As in Lemma \ref{lem:twodimHtwo}, we have a natural identification $\h^* \rightarrow H^2(Y;\C)$. Therefore, the formally universal Poisson deformation is a flat family $\mc{X} \rightarrow \h^* / W$.
Fix a finite group $\Gamma \subset \widetilde{\Xi} \subset N_{SL(2,\C)}(\Gamma)$. Lemma \ref{lem:twodimHtwo} implies that the quotient $\Xi := \widetilde{\Xi} / \Gamma$ acts on $\h^*$ via Dynkin diagram automorphisms. In this case, $\mf{c}$ is the space of all $\Gamma$-equivariant functions $\Gamma \smallsetminus \{ 1 \} \rightarrow \C$, the action of $\Gamma$ on $\C$ being trivial. The group $\Xi$ acts on $\mf{c}$ by $(x \cdot \chi)(s) = \chi(\tilde{x} s \tilde{x}^{-1})$, where $\tilde{x}$ is some lift of $x$ to $\widetilde{\Xi}$. This action extends uniquely to an action of $ \widetilde{\Xi}$ on $\mathbf{H}(\Gamma)$ by algebra automorphisms such that the restriction of this action to $\Gamma$ is just conjugation. The action preserves the spherical subalgebra $e \mathbf{H}(\Gamma) e$, the action of $ \widetilde{\Xi}$ on this subalgebra factoring through $\Xi$. Thus, $\Xi$ acts on $\CM(\Gamma)$ such that the map $\CM(\Gamma) \rightarrow \mf{c}$ is equivariant. Since the action of $\Xi$ on $e \mathbf{H}(\Gamma) e$ can be extended to the case where $t = 1$ (or more generally a formal variable $t$), $\Xi$ acts on $\CM(\Gamma)$ via Poisson automorphisms. Recall that we have defined in (\ref{eq:mcKay}) an isomorphism $\varpi : \mf{c} {\;\stackrel{_\sim}{\to}\;} \h^*$; this is an $\Xi$-equivariant isomorphism, where $\Xi$ acts on $\h^*$ as defined in section \ref{sec:McKay}.
\begin{lem}\label{lem:rank2}
The map $\varpi$ extends to a $\Cs$-equivariant isomorphism $\CM(\Gamma) \simeq \mf{c} \times_{\h^* / W} \mc{X}$.
\end{lem}
\begin{proof}
Let $\mf{g}$ be the simple Lie algebra associated to $\Gamma$ under the McKay correspondence. It is well-known, e.g. \cite[Proposition 3.1 (1)]{Namikawa}, that a Slodowy slice $S \rightarrow \h^* / W$ to the subregular nilpotent orbit in $\mf{g}$ is the formally universal Poisson deformation of the Kleinian singularity $V / \Gamma$. Let $\widetilde{S} \rightarrow \h^*$ be the resolution of the formally universal deformation $S \rightarrow \h^* / W$ of $V / \Gamma$ coming from taking the preimage of $S$ in Grothendieck's simultaneous resolution of $\g^*$. By \cite[Proposition 6.2]{SlodowyNamLehnSorger}, $\widetilde{S} \rightarrow \h^*$ is the formally universal Poisson deformation of the minimal resolution of $V / \Gamma$.
Theorems 6.2.2 and 5.3.1 of \cite{Losev} imply that there is an isomorphism $\CM(\Gamma) {\;\stackrel{_\sim}{\to}\;} \widetilde{S}^{\aff}$ such that the following diagram commutes
$$
\begin{tikzpicture}
\node (C1) at (-1.2,1.2) {$\CM(\Gamma)$};
\node (C2) at (1.2,1.2) {$\widetilde{S}^{\aff}$};
\node (C3) at (-1.2,0) {$\mf{c}$};
\node (C4) at (1.2,0) {$\h^*$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C3) -- (C4);
\node [above] at (0,0) {$\varpi$};
\node [above] at (0,1.2) {$\sim$};
\end{tikzpicture}
$$
This implies the statement of the lemma.
\end{proof}
\subsection{Factorization of the Calogero-Moser space}\label{sec:factor}
Fix $B \in \mc{B}$, $\uGamma$ the corresponding minimal parabolic subgroup and $\mc{L}$ the symplectic leaf in $V / \Gamma$ labeled by $B$. Let $\mf{d} \simeq \h^*_B$ be the base of the Calogero-Moser deformation of $\uGamma$ so that $\mathbf{H}(\uGamma)$ is a $\C[\mf{d}]$-algebra. For clarity, we write $\mathbf{H}_{\mf{d}}(\uGamma) := \mathbf{H}(\uGamma)$ to show the dependence on $\mf{d}$. As explained in section \ref{sec:SRA}, the space $\mf{c}_B = \mf{d}^{\Xi(B)}$ is a subspace and projection followed by inclusion defines a linear map $\mf{c} \rightarrow \mf{d}$. Let $\mathbf{H}_{\mf{c}}(\uGamma) := \C[\mf{c}] \otimes_{\C[\mf{d}]} \mathbf{H}_{\mf{d}}(\uGamma)$ denote the symplectic reflection algebra obtained from $\mathbf{H}(\uGamma)$ by base change from $\mf{d}$ to $\mf{c}$. Choose $p \in \mc{L} \subset V / \Gamma$. We may think of $p$ as a $\Gamma$-orbit in $V$. If $I_p$ is the ideal of functions in $\C[V]$ vanishing on this orbit, then $I_p \rtimes \Gamma$ is a two-sided ideal in $\C[V] \rtimes \Gamma$. Recall that $\C[V] \rtimes \Gamma$ is the quotient of $\mathbf{H}(\Gamma)$ by the ideal generated by $\C[\mf{c}]_+$. Following Losev, we denote by $\mathbf{H}(\Gamma)^{\wedge p}$ the completion of $\mathbf{H}(\Gamma)$ by the preimage of the $I_p \rtimes \Gamma$ under the quotient map. Since the preimage of $I_p \rtimes \Gamma$ in $\mathbf{H}(\Gamma)$ contains $\C[\mf{c}]_+$, the completion $\mathbf{H}(\Gamma)^{\wedge p}$ is a topological $\C[[\mf{c}]]$-algebra. Similarly, $ \mathbf{H}_{\mf{c}}(\uGamma)^{\wedge 0}$ is the completion of $ \mathbf{H}_{\mf{c}}(\uGamma)$ corresponding to the ideal $\C[V_0]_+ \rtimes \uGamma$ of $\C[V_0] \rtimes \uGamma$. The key result \cite[Theorem 2.5.3]{LosevSRAComplete} says
\begin{thm}\label{thm:Losev}
There is an isomorphism
$$
\theta^* : \mathbf{H}(\Gamma)^{\wedge p} \rightarrow \mathrm{Mat}_{|\Gamma / \uGamma|} \left( \mathbf{H}_{\mf{c}}(\uGamma)^{\wedge 0} \widehat{\o} \C[V^{\uGamma}] \right)
$$
of topological $\C[[\mf{c}]]$-algebras.
\end{thm}
Let $e$ and ${e'}$ denote the trivial idempotents in the group algebras of $\Gamma$ and $\uGamma$ respectively, so that $\mathsf{CM}_{\mf{c}}(\uGamma) = \mathop{\rm Spec}\nolimits {e'} \mathbf{H}_{\mf{c}}(\uGamma) {e'} $ is a Poisson variety over $\mf{c}$. Applying the idempotent $e$ to both sides of the isomorphism $\theta^*$ of Theorem \ref{thm:Losev} gives an isomorphism $e (\mathbf{H}(\Gamma)^{\wedge p}) e \rightarrow {e'} ( \mathbf{H}_{\mf{c}}(\uGamma)^{\wedge 0}) {e'} \ \widehat{\o} \ \C[V^{\uGamma}]$; see \cite[Section 2.3]{LosevSRAComplete}. The isomorphism $\theta^*$ of Theorem \ref{thm:Losev} is actually valid for any $t$. This implies that the isomorphism $e( \mathbf{H}(\Gamma)^{\wedge p}) e \rightarrow {e'} ( \mathbf{H}_{\mf{c}}(\uGamma)^{\wedge 0}) {e'} \ \widehat{\o} \ \C[V^{\uGamma}]$ is an isomorphism of \textit{Poisson} algebras.
\begin{cor}\label{cor:Losev}
There is an isomorphism of formal Poisson schemes
$$
\theta : \CM(\Gamma)^{\wedge p} \rightarrow \mathsf{CM}_{\mf{c}}(\uGamma)^{\wedge 0} \times V^{\uGamma}
$$
over $\widehat{\mf{c}}$.
\end{cor}
\subsection{} Recall that Lemma \ref{lem:classmap} says that there is a $\Cs$-equivariant morphism $\alpha : \mf{c} \rightarrow H^2(Y;\C) / W$ such that $\CM(\Gamma) \simeq \mf{c} \times_{H^2(Y;\C) / W} \mc{X}$. Completing at $0 \in \mf{c}$ and $o \in H^2(Y,\C) / W$, we have a Cartesian square
$$
\begin{tikzpicture}
\node (C1) at (-1.5,1.5) {$\CM(\Gamma)^{\wedge}$};
\node (C2) at (1.5,1.5) {$\mc{X}^{\wedge}$};
\node (C3) at (-1.5,0) {$\widehat{\mf{c}}$};
\node (C4) at (1.5,0) {$(H^2(Y;\C) / W)^{\wedge}$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C3) -- (C4);
\node [above] at (-0.6,0) {$\hat{\alpha}$};
\end{tikzpicture}
$$
such that $\alpha$ is the algebraization of $\hat{\alpha}$. Here $\CM(\Gamma)^{\wedge}$ and $\mc{X}^{\wedge}$ are the completions of $\CM(\Gamma)$ and $\mc{X}$ respectively along the special fiber $V / \Gamma$. Choose some $B \in \mc{B}$ and consider $\uGamma := \Gamma_B$, $\mc{L}_B$ etc. Fix $p \in \mc{L}_B$. Completing $\CM(\Gamma)^{\wedge}$ at $p$, the above diagram becomes the Cartesian square
\begin{equation}\label{eq:carA}
\begin{tikzpicture}
\node (C1) at (-1.5,1.5) {$\CM(\Gamma)^{\wedge p}$};
\node (C2) at (1.5,1.5) {$\mc{X}^{\wedge p}$};
\node (C3) at (-1.5,0) {$\widehat{\mf{c}}$};
\node (C4) at (1.5,0) {$(H^2(Y;\C) / W)^{\wedge}$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C3) -- (C4);
\node [above] at (-0.6,0) {$\hat{\alpha}$};
\end{tikzpicture}
\end{equation}
Recall that Theorem \ref{prop:Weylgroup} says that $H^2(Y;\C) / W$ is isomorphic to $\prod_{B \in \mc{B}} \mf{c}_B / W_B$. The projection map $H^2(Y;\C) / W \rightarrow \mf{c}_B / W_B$, followed by the canonical morphism $\mf{c}_B / W_B \rightarrow \h_B^* / W(B)$ is denoted $r_B$, and its completion at $0$ is $\hat{r}_B$. Let $\mc{X}_0(\uGamma)$ denote the formally universal Poisson deformation of $V_0 / \uGamma$. The completion of $\mc{X}_0(\uGamma)$ at $o \in V_0 / \uGamma \subset \mc{X}_0(\uGamma)$ is denoted $\mc{X}_0(\uGamma)^{\wedge 0}$.
\begin{lem}\label{lem:Q2}
Let $p \in \mc{L}_B$ and $\mc{X}^{\wedge p}$ the completion of $\mc{X}$ at $p$. Then the following commutative diagram
\begin{equation}\label{eq:carB}
\begin{tikzpicture}
\node (C1) at (-2,1.5) {$\mc{X}^{\wedge p}$};
\node (C2) at (2,1.5) {$\mc{X}_0(\uGamma)^{\wedge 0} \times (V^{\uGamma})^{\wedge 0}$};
\node (C3) at (-2,0) {$(H^2(Y;\C) / W)^{\wedge}$};
\node (C4) at (2,0) {$(\h_B^* / W(B))^{\wedge}$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C3) -- (C4);
\node [above] at (0.2,0) {$\hat{r}_B$};
\end{tikzpicture}
\end{equation}
is Cartesian.
\end{lem}
\begin{proof}
The analytic germ of $0$ in $H^2(Y,\C) / W$, resp. in $\h_B^* / W(B)$, is denoted $\mathrm{PDef}(V / \Gamma)$, resp. $\mathrm{PDef}(V / \uGamma)$. They are the Poisson Kuranishi spaces of the corresponding analytic symplectic varieties. Passing to the analytic topology, the formally universal deformation $\mc{X} \rightarrow H^2(Y;\C) / W$ induces a flat Poisson deformation $\mc{X}^{\an} \rightarrow (H^2(Y;\C) / W)^{\an}$. Restricting to the germ of $o$ in $H^2(Y;\C) / W$, we have a flat Poisson deformation $\mc{X}^{\an} \rightarrow \mathrm{PDef}(V / \Gamma)$.
Passing to the germ of $p \in (V / \Gamma)^{\an} \subset \mc{X}^{\an}$ gives a flat family $(\mc{X}^{\an}, p) \rightarrow \mathrm{PDef}(V / \Gamma)$. This is a deformation of $((V / \Gamma)^{\an}, p)$. By the generalized Darboux Theorem, \cite[Lemma 1.3]{Namikawa}, we have an isomorphism of symplectic varieties
$$
((V / \Gamma)^{\an}, p) \simeq ((V_0 / \uGamma \times V^{\uGamma})^{\an}, 0).
$$
Moreover, by \cite[Proposition 3.1]{Namikawa}, the universal Poisson deformation of $((V_0 / \uGamma)^{\an}, 0) \times ( (V^{\uGamma})^{\an}, 0)$ is $((\mc{X}_0(\uGamma) \times V^{\uGamma})^{\an}, 0) \rightarrow \mathrm{PDef}(V_0 / \uGamma)$. Hence there exists a holomorphic map $\phi_B : \mathrm{PDef}(V / \Gamma) \rightarrow \mathrm{PDef}(V_0 / \uGamma)$ such that the following diagram is Cartesian
$$
\begin{tikzpicture}
\node (C1) at (-2,1.5) {$(\mc{X}^{\an}, p)$};
\node (C2) at (2,1.5) {$((\mc{X}_0(\uGamma) \times V^{\uGamma})^{\an}, 0)$};
\node (C3) at (-2,0) {$\mathrm{PDef}(V / \Gamma)$};
\node (C4) at (2,0) {$ \mathrm{PDef}(V_0 / \uGamma)$};
\draw[->] (C1) -- (C2);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C3) -- (C4);
\node [above] at (0,0) {$\phi_B$};
\end{tikzpicture}
$$
The map $\phi_B$ is precisely the map constructed in section 4 (i) of the proof of \cite[Theorem 1.1]{Namikawa2}. As explained in section 4 of the proof of of \cite[Theorem 1.1]{Namikawa2}, the completion of $\phi_B$ equals $\hat{r}_B$. Passing to the formal neighborhood of $p$ in $(\mc{X}^{\an}, p)$ and $0$ in $((\mc{X}_0(\uGamma) \times V^{\uGamma})^{\an}, 0)$, we get the Cartesian square stated in the lemma.
\end{proof}
\subsection{The proof of Theorem \ref{thm:main}} By Corollary \ref{cor:Losev}, we have an isomorphism of Poisson varieties $\CM(\Gamma)^{\wedge p} \simeq \mathsf{CM}_{\mf{c}}(\uGamma)^{\wedge 0} \times V^{\uGamma} $ over $\widehat{\mf{c}}$. Under the identification (\ref{eq:mcKay}), we have a natural decomposition $\mf{c} = \bigoplus_{B \in \mc{B}} \mf{c}_B$. Therefore, if we write $q_B$ for projection from $\mf{c}$ onto $\mf{c}_B$, the square
\begin{equation}\label{eq:carC}
\begin{tikzpicture}
\node (C1) at (-1.5,1.5) {$\mathsf{CM}_{\mf{c}}(\uGamma)^{\wedge 0} \times V^{\uGamma}$};
\node (C2) at (1.5,1.5) {$\CM(\Gamma)^{\wedge p}$};
\node (C3) at (-1.5,0) {$\widehat{\mf{c}}_B$};
\node (C4) at (1.5,0) {$\widehat{\mf{c}}$};
\draw[->] (C2) -- (C1);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C4) -- (C3);
\node [above] at (0,0) {$\hat{q}_B$};
\end{tikzpicture}
\end{equation}
is also (trivially) Cartesian. Recall from section \ref{sec:factor} that $\mf{d}$ is the natural parameter space associated to the symplectic reflection algebra $\mathbf{H}_{\mf{d}}(\uGamma)$ and $\mf{c}_B = \mf{d}^{\Xi(B)}$. Let $\alpha_B$ be the composite $\mf{c}_B \hookrightarrow \mf{d} {\;\stackrel{_\sim}{\to}\;} \h^*_B \rightarrow \h^*_B / W(B)$ and $\hat{\alpha}_B$ will be its completion at $0$. Lemma \ref{lem:rank2} implies that the following diagram is Cartesian
\begin{equation}\label{eq:carD}
\begin{tikzpicture}
\node (C1) at (-1.5,1.5) {$\mc{X}_0(\uGamma)^{\wedge 0} \times V^{\uGamma}$};
\node (C2) at (1.5,1.5) {$\CM_{\mf{c}}(\uGamma)^{\wedge 0} \times V^{\uGamma}$};
\node (C3) at (-1.5,0) {$(\h_B / W(B))^{\wedge}$};
\node (C4) at (1.5,0) {$ \widehat{\mf{c}}_B$};
\draw[->] (C2) -- (C1);
\draw[->] (C1) -- (C3);
\draw[->] (C2) -- (C4);
\draw[->] (C4) -- (C3);
\node [above] at (0.5,0) {$\hat{\alpha}_B$};
\end{tikzpicture}
\end{equation}
The composite of the two bottom horizontal arrows is denoted $\hat{\alpha}_B$. By Lemma \ref{lem:rank2}, $\hat{\alpha}_B = \hat{\beta}_B$. Combining diagrams (\ref{eq:carA}), (\ref{eq:carB}), (\ref{eq:carC}) and (\ref{eq:carD}), we get the following diagram, where each square is Cartesian.
$$
\begin{tikzpicture}
\node (C1) at (-6,1.5) {$\mc{X}_0(\uGamma)^{\wedge} \times V^{\uGamma}$};
\node (C2) at (-3,1.5) {$\CM_{\mf{c}}(\uGamma)^{\wedge 0} \times V^{\uGamma}$};
\node (C3) at (0,1.5) {$\CM(\Gamma)^{\wedge p}$};
\node (C4) at (2.5,1.5) {$\mc{X}^{\wedge p}$};
\node (C5) at (5.7,1.5) {$\mc{X}_0(\uGamma)^{\wedge 0} \times V^{\uGamma}$};
\node (C6) at (-6,0) {$(\h_B / W(B))^{\wedge}$};
\node (C7) at (-3,0) {$\widehat{\mf{c}}_B$};
\node (C8) at (0,0) {$\widehat{\mf{c}}$};
\node (C9) at (2.5,0) {$(H^2(Y,\C) / W)^{\wedge}$};
\node (C10) at (5.7,0) {$(\h_B^* / W(B) )^{\wedge}$};
\draw[->] (C2) -- (C1);
\draw[->] (C3) -- (C2);
\draw[->] (C3) -- (C4);
\draw[->] (C4) -- (C5);
\draw[->] (C1) -- (C6);
\draw[->] (C2) -- (C7);
\draw[->] (C3) -- (C8);
\draw[->] (C4) -- (C9);
\draw[->] (C5) -- (C10);
\draw[->] (C7) -- (C6);
\draw[->] (C8) -- (C7);
\draw[->] (C8) -- (C9);
\draw[->] (C9) -- (C10);
\node [above] at (-4,0) {$\hat{\alpha}_B$};
\node [above] at (-1.4,0) {$\hat{q}_B$};
\node [above] at (0.7,0) {$\hat{\alpha}$};
\node [above] at (4.25,0) {$\hat{r}_B$};
\end{tikzpicture}
$$
The universality of the formal Poisson deformation $\mc{X}_0(\uGamma)^{\wedge} \times V^{\uGamma} \rightarrow (\h_B^* / W(B))^{\wedge}$ of $V / \uGamma = V_0/ \uGamma \times V^{\uGamma}$ implies that $\hat{\alpha}_B \circ \hat{q}_B = \hat{p}_B \circ \hat{\alpha}$, c.f. \cite[Section 1.3]{GK}. Hence
$$
\bigoplus_{B \in \mc{B}} \hat{\alpha}_B \circ \hat{q}_B = \bigoplus_{B \in \mc{B}} \hat{r}_B \circ \hat{\alpha} = \hat{\alpha}
$$
since $\bigoplus_{B \in \mc{B}} \hat{r}_B = \mathrm{id}$. On the other hand, it is clear from the explicit definition of $\hat{\beta}$ that $\hat{\beta} = \bigoplus_{B \in \mc{B}} \hat{\alpha}_B \circ \hat{q}_B$. This completes the proof of Theorem \ref{thm:main}.
\subsection{}\label{sec:BR}
We turn to the proof of Corollary \ref{cor:singCal}. If the quotient $V / \Gamma$ admits a smooth projective, symplectic resolution then by Theorem \ref{thm:main} and the main theorem of \cite{Namikawa3}, the set of points in $\mf{c}$ for which $\CM_{\bc}(\Gamma)$ is singular is a union of hyperplanes. If $V / \Gamma$ does not admit a symplectic resolution then, by \cite[Corollary 1.21]{GK}, the space $\CM_{\bc}(\Gamma)$ is always singular.
\section{Counting resolutions}
In this section we deduce Theorem \ref{thm:count} from Theorem \ref{thm:main}, using the main theorem from \cite{Namikawa3}.
\subsection{}\label{sec:proof1} We recall the main result from \cite{Namikawa3}. As explained in \textit{loc. cit.}, the birational geometry of $Y$ is controlled by the real space $H^2(Y;\R)$. In particular, a key role is played by the movable cone $\mathrm{Mov}(\rho) \subset H^2(Y;\R)$ of $\rho : Y \rightarrow V / \Gamma$; see \textit{loc. cit.} for the definition. The ample cone of $\rho$ in $H^2(Y;\C)$ is denoted $\mathrm{Amp}(\rho)$. By Theorem \ref{thm:main}, we have a projective morphism $\mc{Y} \rightarrow \CM(\Gamma)$ over $\mf{c} \simeq H^2(Y;\C)$. As explained in \cite{Namikawa3}, the set $\mc{D} \subset \mf{c}$ defined in the introduction corresponds to the closed subset $\mc{D}' \subset H^2(Y;\C)$ consisting of all points $t$ such that the fiber $\mc{Y}_t := \boldsymbol{\nu}^{-1}(t)$ is not affine.
\begin{thm}[Main Theorem, \cite{Namikawa3}]\label{thm:Nam3main}
\begin{enumerate}
\item There are finitely many hyperplanes $\{H_i \}_{i \in I}$ in $H^2(Y;\Q)$ such that $\mc{D}' = \bigcup_{i \in I} (H_i)_{\C}$.
\item There are only finitely many $\Q$-factorial terminalizations $\{ \rho_k : Y_k \rightarrow V / \Gamma \}_{k \in K}$ of $V / \Gamma$.
\item The set of open chambers determined by the real hyperplanes $\{ (H_i)_{\R} \}_{i \in I}$ coincides with the set $\{ w(\mathrm{Amp}(\rho_k)) \}$, where $w \in W$ and $k \in K$.
\end{enumerate}
\end{thm}
For a topological space $X$, we abuse notation and let $\pi_0(X)$ denote the number of connected components of $X$. Let $\mathrm{Mov}(\rho)^{\circ} = \mathrm{Mov}(\rho) \smallsetminus \bigcup_{i \in I} (H_{i})_{\R}$. By Theorem \ref{thm:Nam3main} (3) and \cite[Lemma 6]{Namikawa3}, $\mathrm{Mov}(\rho)^{\circ}$ equals $\bigsqcup_{k \in K} \mathrm{Amp}(\rho_k)$. Hence, the number $|K|$ of pairwise non-isomorphic $\Q$-factorial terminalizations of $V / \Gamma$ equals $\pi_0(\mathrm{Mov}(\rho)^{\circ})$. Proposition 2.19 of \cite{BPWQuant1} implies that
$$
H^2(Y;\R) \smallsetminus \bigcup_{i \in I} (H_{i})_{\R} = \bigsqcup_{w \in W} w(\mathrm{Mov}(\rho)^{\circ}).
$$
Thus,
\begin{equation}\label{eq:pizero}
\pi_0 \left( H^2(Y; \R) \smallsetminus \bigcup_i H_i \right) = |W| \cdot |K|.
\end{equation}
Zaslavsky's Theorem, \cite[Theorem 2.68]{OrlikTeraoBook}, says that the number of connected components of the complement $H^2(Y; \R) \smallsetminus \bigcup_i (H_i)_{\R}$ to the real hyperplane arrangement $\bigcup_i (H_i)_{\R}$ equals the dimension of the cohomology ring of the complement $H^2(Y; \C) \smallsetminus \mc{D}'$ to the complexification $\mc{D}'$ of the real hyperplane arrangement. As explained above, Theorem \ref{thm:main} and Theorem \ref{thm:Nam3main} (1) imply that $H^2(Y; \C) \smallsetminus \mc{D}' \simeq \mf{c} \smallsetminus \mc{D}$. Hence
\begin{equation}\label{eq:pizeroC}
\pi_0 \left( H^2(Y; \R) \smallsetminus \bigcup_i H_i \right) = \dim_{\C} H^*(\mf{c} \smallsetminus \mc{D}; \C).
\end{equation}
Thus, the claim of Theorem \ref{thm:count} follows from equations (\ref{eq:pizero}) and (\ref{eq:pizeroC}).
\subsection{The wreath product $\s_n \wr G$.}\label{sec:wreath} In this section we deduce Proposition \ref{prop:countWeyl} from Theorem \ref{thm:count}. The proposition is trivially true for $n = 1$ due the uniqueness of minimal resolutions. Therefore we assume that $n > 1$. In this case, the Namikawa Weyl group $W$ of $V / \Gamma$ equals $\Z_2 \times W_G$. Let $\h$ be the Cartan algebra on which $W_G$ acts. By \cite[Theorem 1.4]{MarsdenWeinsteinStratification}, there is an isomorphism of vector spaces $\mf{c} \simeq \C \times \h$, lifting to an identification of $\CM(\Gamma)$ with a certain moduli space of representations of a deformed preprojective algebra - see \textit{loc. cit.} for details. The proof of Lemmata 4.3 and 4.4 of \cite{GordonQuiver} shows that the set $\mc{D}$ over which the fibers $\CM_{\bc}(\Gamma)$ are singular is the union of hyperplanes in $\C \times \h$ given by
$$
H_{\lambda,m} := \{ (\alpha, x) \in \C \times \h \ | \ \lambda(x) + m \alpha = 0 \} \quad \textrm{and} \quad \alpha = 0,
$$
where $\lambda \in R$, the root system of the Weyl group $W_G$, and $1-n \le m \le n-1$. In the language of \cite[Chapter 1]{OrlikTeraoBook}, this hyperplane arrangement is the cone over the affine hyperplane arrangement
$$
\mc{A}= \left\{ \overline{H}_{\lambda,m} = \{ x \in \h \ | \ \lambda(x) + m = 0 \} \ | \ \lambda \in R, \ 1-n \le m \le n-1 \ \right\}.
$$
Therefore \cite[Proposition 2.51]{OrlikTeraoBook} implies that
$$
\dim_{\C} H^*(\mf{c} \smallsetminus \mc{D}; \C) = 2 \dim_{\C} H^* \left(\h \smallsetminus \bigcup_{H \in \mc{A}} H; \C\right).
$$
Since $|W| = 2 |W_G|$, the result follows from the above equality, together with equation (1) of Theorem 1.2 in \cite{athanasiadis} and the fact that $|W_G| = \prod_{i = 1}^{\ell} (e_i + 1)$.
\subsection{Exceptional groups}\label{sec:except} Other than the infinite series $\s_n \wr G$, there are only two exceptional groups that are known to admit symplectic resolutions. These are $Q_8 \times_{\Z_2} D_8$ and $G_4$, and their explicit descriptions as subgroups of $\mathrm{Sp}(\C^4)$ can be found in \cite{smoothsra} and \cite{Singular} respectively.
First we consider the group $Q_8 \times_{\Z_2} D_8$. As shown in \cite{smoothsra}, we have $\mf{c} = \C \{ \bc_1, \ds, \bc_5 \}$ and $\mc{B}$ consists of five elements $B_1, \ds, B_5$. Each minimal parabolic $\underline{\Gamma}_{B_i}$ is isomorphic to $\Z_2$ and the corresponding quotients $\Xi(B)$ are always trivial. Thus, $W_{B_i} = \s_2$ for all $i$ and if $a_i$ is the generator of $W_{B_i}$ then $a_i \cdot \bc_j = (-1)^{\delta_{ij}} \bc_j$. There are 21 hyperplanes in $\mf{c}$ given by the 16 of the form $\bc_1 \pm \bc_2 \pm \bc_3 \pm \bc_4 \pm \bc_5 = 0$ and the five of the from $\bc_i = 0$. Using\footnote{A copy of the code used to make this calculation is available from the author.} the computer algebra program $\mathrm{MAGMA}$ \cite{MAGMA}, it is possible to calculate that the Poincar\'e polynomial of the Orlik-Solomon algebra equals $1 + 21 t + 170 t^2 + 650 t^3 + 1125 t^4 + 625 t^5$. This implies that the quotient $V / \Gamma$ admits $81$ distinct symplectic resolutions.
For the group $G_4$, the proof of \cite[Theorem 1.4]{MoPositiveChar} shows that $\mc{D} = H_1 \cup H_2 \cup H_3$, where
$$
H_1 = \{ \bc_1 + \bc_2 = 0 \}, \quad H_2 = \{ \omega \bc_1 + \omega ^2 \bc_2 = 0 \}, \quad H_3 = \{ \omega^2 \bc_1 + \omega \bc_2 = 0 \},
$$
with $\omega$ a primitive $3$rd root of unity. Since $\dim H_i \cap H_j = 0$ for $i \neq j$, the only dependent subset of $\{ H_1, H_2, H_3 \}$ is $\{ H_1, H_2, H_3 \}$ itself. Therefore, \cite[Definition 3.5]{OrlikTeraoBook} says that the Orlik-Solomon algebra associated to the arrangement $\mc{D}$ is the quotient of the exterior algebra $\wedge^{{\:\raisebox{2pt}{\text{\circle*{1.5}}}}}(x_1,x_2,x_3)$ by the two-sided ideal generated by
$$
\partial (x_1 \wedge x_2 \wedge x_3) = x_2 \wedge x_3 - x_1 \wedge x_3 + x_1 \wedge x_2.
$$
Hence, the Orlik-Solomon Theorem \cite[Theorem 5.90]{OrlikTeraoBook} says that
$$
H^{{\:\raisebox{2pt}{\text{\circle*{1.5}}}}}(\mf{c} \smallsetminus \mc{D},\C) \simeq \frac{\wedge^{{\:\raisebox{2pt}{\text{\circle*{1.5}}}}}(x_1,x_2,x_3)}{\langle x_2 \wedge x_3 - x_1 \wedge x_3 + x_1 \wedge x_2 \rangle}.
$$
The Orlik-Solomon algebra has basis $\{ 1, x_1, x_2, x_3, x_1 \wedge x_3 ,x_1 \wedge x_2 \}$; this can be seen directly, or by applying \cite[Theorem 3.43]{OrlikTeraoBook}. The Weyl group for $G_4$ is $\Z_3$. Hence Theorem \ref{thm:count} implies that there are $2$ non-isomorphic symplectic resolutions of $\C^4 / G_4$. This implies that the two symplectic resolutions constructed in \cite{LehnSorger} exhaust all symplectic resolutions.
\def$'${$'$} \def$'${$'$} \def$'${$'$} \def$'${$'$}
\def$'${$'$} \def$'${$'$} \def$'${$'$} \def$'${$'$}
\def$'${$'$} \def$'${$'$} \def$'${$'$} \def$'${$'$}
\def$'${$'$}
|
\section{Introduction}
A Bose-Einstein condensate of a dilute gas of alkaline atoms in a double well potential realizes the physics of Josephson junctions, which was originally predicted in two superconductors separated by an insulating layer \cite{josephson1962possible}. The bosonic realization of Josephson junction physics has attracted a great interest both theoretically \cite{javanainen1986oscillatory, smerzi1997quantum, raghavan1999coherent, milburn1997quantum, mazzarella2010spontaneous, mazzarella2011coherence, julia2010macroscopic, julia2010bose, gillet2014tunneling} and experimentally \cite{albiez2005direct, levy2007ac, zibold2010classical} in recent years. On one hand the physics of Josephson junctions can be described by the two coupled nonlinear equations of a non-rigid pendulum, therefore its careful investigation is very tempting, since the model and its mathematics look fairly simple, while they are complicated enough in order to help us understanding some aspects of more elaborate problems, like the Bose-Hubbard model. In particular, bosonic Josephson junctions (BJJs) may be regarded as a two-site realization of the Bose-Hubbard model. On the other hand the mesoscopic coherent dynamics of Bose-Einstein condensate has important issues of its own, such as the validity of semiclassical dynamics and the use of coherent states in few mode and finite atom number systems \cite{mazzarella2011coherence, julia2010macroscopic}.
The tunneling dynamics of BJJs can serve as a basic tool in interferometry applications \cite{shin2004atom, shin2005interference, schumm2005matter}. The first experiments with repulsively interacting Bose condensates revealed self-trapping and plasma oscillations \cite{albiez2005direct} and later, with an experimental effort the a.c. Josephson effect was also observed \cite{levy2007ac}. With the help of atomic Feshbach resonances it is possible to change the magnitude and even the sign of the parameter of on-site interaction. Therefore it is in principle possible to ``quench'' the dynamics of the BJJ and realize the semiclassical dynamics around the stationary points of the Josephson equations or change the dynamics governed by one particular fixed point to a different one governed by a different fixed point \cite{zibold2010classical}. This way a setup for very fast macroscopic entanglement generation can be achieved \cite{micheli2003many,Vidal2004Entanglement}.
The question naturally arises whether it is possible or not to obtain some similar quenching not only with the on-site interaction, but rather by engineering the tunneling amplitude of the junction? In this paper we give an affirmative answer to this question. With the help of an external, tightly focused, red-detuned laser beam one can create a tiny hole in the middle of the double-well barrier. When the depth of this dip is increased, at some point a bound state localized inside the dip potential appears, and by further increasing the potential depth the tunneling constant between the original left and right wells changes sign. The creation of such a static obstacle is fairly simple and therefore gives another knob on the system besides the standard Feshbach resonance technique.
The plan of the paper is as follows. In Sec. \ref{sec:dwd} we consider the single particle problem where a dip potential is superimposed on the standard double well. In Sec. \ref{sec:bjj} we apply the two-mode approximation to the problem when the doublet formed by the Wannier states of the left and right wells are sufficiently separated from the other energy levels and consider the Josephson dynamics. We summarize in Sec. \ref{sec:sum}. The stability analysis of the stationary points of the dynamics is moved to the Appendix.
\section{Double well with a dip in the middle}
\label{sec:dwd}
\begin{figure*}[ht!]
\centering
\includegraphics[width=\textwidth]{LowestStates}
\caption{(Color online) Top panel: the potential landscape with the energy eigenvalues for various laser intensities $I_0$. Bottom panel: the wave functions corresponding to the lowest three energies. $v_1(x)$ is the ground state wave function, $v_2(x)$ is the wave function of the first excited state, and $v_3(x)$ is the second excited state. From left to right the parameter $I_0$ varies as: $0.0, 2.0, 4.0, 8.0$}
\label{fig:lowest_states}
\end{figure*}
The double well setup considered here consists of a symmetric potential
\begin{equation}
\label{eq:dwpot}
V_{\text{DW}}(x)=\frac{1}{2}m\omega_H^2 x^2+V_1\, e^{-\frac{x^2}{2 w^2}},
\end{equation}
where $m$ is the mass of the atoms, $\omega_H$ is the frequency of the parabolic confinement, $V_1$ is the height and $w$ is the width of the double well barrier. We consider tight confinement in the perpendicular directions and treat the system as one-dimensional. In addition to the double well potential there is a tightly focused laser beam in the center which is red detuned from the atomic transition creating a further attractive potential for the atoms,
\begin{equation}
\label{eq:centerwell}
V_{L}(x)=-I_0\,e^{-\frac{x^2}{2 \sigma^2}}
\end{equation}
where $I_0$ is the strength and $\sigma\ll w$ is the width of the optical potential. The full single particle Hamiltonian is
\begin{equation}
\label{eq:ham}
\hat H=-\frac{\hbar^2}{2m}\frac{d^2}{dx^2} + V_{\text{DW}}(x)+V_L(x).
\end{equation}
The perturbing potential $V_L(x)$ opens up a narrow dip in the center of the double well barrier, as illustrated in Fig. \ref{fig:lowest_states}. When varying the strength $I_0$, one can interpolate between a symmetric double well potential and a triple well one. For $I_0=0$, with our choice of parameters (for ${}^{87}\mathrm{Rb}$), which is close to experimental applications ($m=87\,\mathrm{amu}$, $\omega_H=2\pi\times15\,\mathrm{Hz}$, $w=5\,\mathrm{\mu m}$, $V_1=5 m \omega_H^2 w^2$, and $\sigma=0.5\,\mathrm{\mu m}$) the lowest two energy eigenvalues are almost degenerate and they form the low energy doublet of the double well problem. The corresponding wave functions are the symmetric and antisymmetric combinations of the Wannier orbits, which themselves are localized states around the left and right energy minima of the potential. Other energy eigenvalues are much higher and one can rely on a two-mode approximation when treating the problem.
\begin{figure}[b!]
\centering
\includegraphics[width=\columnwidth]{energy_crossing}
\caption{(Color online) The three lowest energy eigenvalues plotted as a function of $I_0$. One can observe an avoided crossing. The second energy level is unaffected by the perturbing potential, while the lowest energy and the third energy eigenvalue tilt down with increasing $I_0$.}
\label{fig:energy_crossing}
\end{figure}
When $I_0$ is increased gradually, as shown in the subsequent plots in Fig \ref{fig:lowest_states}, a central well starts to form in the middle of the potential barrier. For small values of $I_0$ the central well doesn't support a localized state and its effect is just a small perturbation of the energy eigenvalues and an even smaller one on the wave functions. The three lowest energy eigenvalues are plotted in Fig. \ref{fig:energy_crossing}. One eigenvalue of the doublet is basically unchanged by the perturbation, namely the one which corresponds to the antisymmetric wave function, which has a node at the position of the perturbation. The other eigenvalue is shifted a little bit downwards. As $I_0$ increases, the central well deepens, and the third energy eigenvalue approaches the low energy doublet. As this third energy eigenvalue comes closer and closer, the two-mode description becomes more and more inaccurate. One can observe an avoided crossing in the three lowest energy eigenvalues. For small values of $I_0$ the lowest two eigenvalues form the doublet of the symmetric and antisymmetric combinations of the Wannier orbits. On the other side of the crossing, i.e. for large values of $I_0$, the single lowest energy eigenvalue correspond to the state localized in the central well, while the next two eigenvalues form now the doublet of the antisymmetric and symmetric combinations of the Wannier orbits localized at the left and right valleys.
\section{Bosonic Josephson Junction}
\label{sec:bjj}
When the splitting of the low energy doublet is much smaller than the energy difference between the doublet and the closest other energy eigenvalue, the two-mode approximation gives a sufficiently accurate description of the tunneling dynamics between the left and right wells. In this limit the other states are non-resonant and energy conservation decouples them from the tunneling dynamics. With the present parameters it means approximately either $I_0<I_{c,1}\approx6$, or $I_0>I_{c,2}\approx7$.
When $I_0<I_{c,1}$ the Wannier functions are given by: $w_1(x)=(v_1(x)+v_2(x))/\sqrt{2}$, $w_2(x)=(v_1(x)-v_2(x))/\sqrt{2}$, for the left and right wells, respectively. For $I_0>I_{c,2}$ the first and second excited states give the Wannier functions, and they read as: $w_1(x)=(v_2(x)+v_3(x))/\sqrt{2}$, and $w_2(x)=(v_2(x)-v_3(x))/\sqrt{2}$, for the left and right wells, respectively. In second quantized form the non-interacting Hamiltonian \eqref{eq:ham} can be cast to the following form:
\begin{equation}
\label{eq:hamfree}
\hat H_0=\epsilon\left(\hat b_1^\dagger \hat b_1+\hat b_2^\dagger \hat b_2\right)-J\left(\hat b_1^\dagger \hat b_2+\hat b_2^\dagger \hat b_1\right),
\end{equation}
where the parameters are given by $\epsilon=\left< w_1\right|\hat H\left| w_1\right>$, and $J=-\left< w_1\right|\hat H\left| w_2\right>$. In the two-mode approximation the total atom number $\hat N=\hat b_1^\dagger \hat b_1+\hat b_2^\dagger \hat b_2$ is a constant of motion, therefore the first term in Eq. \eqref{eq:hamfree} can be dropped. The parameter $J$ shows a ``resonance" like behavior as a function of $I_0$, as illustrated in Fig. \ref{fig:Jres}. We note that the central part of the figure, where the crossing of the energy levels takes place, is not reliable, since the two-mode approximation breaks down. Nevertheless, the tunneling amplitude changes sign at the crossing and the lower energy orbital of the doublet changes from ungerade to gerade symmetry.
\begin{figure}[b!]
\centering
\includegraphics[width=\columnwidth]{Jres}
\caption{(Color online) The tunneling ratio as a function of the depth of the central well, $I_0$.}
\label{fig:Jres}
\end{figure}
\begin{figure*}
\centering
\includegraphics[width=\textwidth]{grid_figure}
\caption{(Color online) Time evolution of the solution of the BJJ equations when a laser dip is abruptly turned on at $t=10$ (time is measured in units of $|J|^{-1}$). For $t<10$ the dip potential is switched off and the parameters are $U N/J=3.5$. The initial conditions are $(z(0)=0.5, \theta(0)=0)$. The system performs Josephson oscillations. At $t=10$ a dip potential is suddenly turned on and kept constant. For $t>10$ the parameters change to $U N/J=-3.5$ and the dynamics exhibits self-trapping. b) shows the population imbalance as a function of time. Subfigure a) shows, for $t<10$, the phase space, fixed points ($\mathbf{X}_1$ and $\mathbf{X}_4$), and the oscillation (thick line) corresponding to the initial conditions. Panel c) shows the phase space for the new system parameters valid from $t=10$, the fixed points $(\mathbf{X}_2$ and $\mathbf{X}_3$). The thick line corresponds to the new trajectory of the system continuing its oscillation in the new energy landscape.}
\label{fig:quench}
\end{figure*}
In the presence of interaction the Hamiltonian is modified to $\hat H=\hat H_0 + \hat H_I$, with
\begin{equation}
\hat H_I=\frac{U}{2}\left(\hat b_1^\dagger\hat b_1^\dagger\hat b_1\hat b_1+\hat b_2^\dagger\hat b_2^\dagger\hat b_2\hat b_2\right),
\end{equation}
where $U$ characterizes the on-site interaction. At sufficiently low temperatures the bosons form a Bose-Einstein condensate, and in the semi-classical approximation the atomic operators are replaced with c-numbers: $b_k=\sqrt{N_k}(t)e^{i\theta_k(t)}$, where $N_k(t)$ is the atom number in well $k$ at time $t$, and $\theta_k(t)$ is the corresponding phase. The total atom number $N_1(t)+N_2(t)\equiv N$ is constant. It is convenient to introduce the fractional population difference of the two wells, $z(t)=[N_1(t)-N_2(t)]/N$, and the relative phase $\theta(t)=\theta_2(t)-\theta_1(t)$. Using this substitution in the Hamiltonian $\hat H$ one can arrive to the semi-classical energy function \cite{smerzi1997quantum}
\begin{equation}
\label{eq:semen}
\mathcal{H}(z,\theta)=-2JN\sqrt{1-z^2}\cos(\theta)+\frac{U}{2}N^2z^2,
\end{equation}
from which the semi-classical equations, known as the bosonic Josephson junction equations can be derived as
\begin{subequations}
\label{eqs:BJJ}
\begin{align}
\dot z&=-\frac{1}{N}\frac{\partial \mathcal{H}}{\partial \theta}=-2J\sqrt{1-z^2}\sin(\theta),\\
\dot\theta&=\frac{1}{N}\frac{\partial \mathcal{H}}{\partial z}=\left(U\,N+\frac{2J}{\sqrt{1-z^2}}\cos(\theta)\right)z.
\end{align}
\end{subequations}
Here and from now on we work with $\hbar=1$. Equations \eqref{eqs:BJJ} have 4 stationary solutions $\dot{\bar{z}}=0$ and $\dot{\bar{\theta}}=0$: It has two zero imbalance solutions with $\mathbf{X}_1=(\bar z=0, \bar \theta=0)$, and $\mathbf{X}_2=(\bar z=0, \bar \theta=\pi)$. Furthermore there are two finite imbalance solutions: $\mathbf{X}_3=(\bar z=\sqrt{1-(2J/U N)^2}, \bar \theta=0)$, and $\mathbf{X}_4=(\bar z=\sqrt{1-(2J/U N)^2}, \bar \theta=\pi)$. By substituting the stationary solutions to the semi-classical energy function \eqref{eq:semen}, one can immediately see, that for $U>0$ the zero imbalance solutions always have the lowest energy. Also depending on the sign of $J$ the minimal energy solution is either with $\bar\theta=0$ for $J>0$, and $\bar\theta=\pi$ for $J<0$. For attractive interaction $U<0$ the finite imbalance solutions are energetically more favorable for $(UN)^2>4J^2$, and the tunneling dynamics exhibits self-trapping \cite{raghavan1999coherent}. Thus, points of $(\bar{z},0)$ with $\bar{z} \neq 0$ are stable fixed points for the ODEs \eqref{eqs:BJJ} only in the presence of attractive on-site interactions $U$ provided that $U<-2|J|/N$. Under inital conditions $(z(0),0)$ - with $z(0)<(2/\Gamma) (\Gamma-1)^{0.5}$ ($\Gamma=|UN/2J|$) - the solutions of these ODEs describes oscillations of the fractional imbalance and relative phase about a nonzero time averaged value and zero, respectively.
By suitably tuning $I_0$, one can change the sign of $J$ by moving from the left side of the resonance to the right side of it (see Fig. \ref{fig:Jres}). All the above condition thus can be satisfied and one can quench between self-trapping and Josephson dynamics (and vice versa), even with repulsive boson-boson interaction. In Fig. \ref{fig:quench} we illustrate the quench dynamics for a repulsive Bose condensate prepared initially for $(z(0)=0.5, \theta(0)=0)$. At $t=0$ the the dip potential is turned off and we have a symmetric double well potential with $J>0$. The system starts Josephson (plasma) oscillations. In panel (a) we show the phase space trajectories and fixed points of the semi-classical Hamiltonian \eqref{eq:semen} for $U N=3.5 J$. The shading corresponds to the energy, where the central (orange) region is the energy minimum and the outer (green) regions correspond to higher energies. The thick line shows the trajectory of the initial Josephson oscillation. On panel (b) we show the population imbalance as a function of time, measured in units of $|J|^{-1}$. The system parameters are left unchanged for $t=10 J^{-1}$. At $t=10 J^{-1}$ we switch on abruptly a dip potential with $I_0$ such to go to the other side of the resonance with $J\rightarrow -J$. Now the phase space diagram is depicted in panel (c). As we see, due to the change of the sign of $J$, the energy landscape changes by $\theta\rightarrow \theta+\pi$, and the finite imbalance (unstable) fixed points corresponding to the energy maxima are moved to the center. The Bose condensate continues its dynamics in the modified landscape, around the $\mathbf{X}_3$ fixed point, which is selected by its instantaneous state $(z(10|J|^{-1}), \theta(10|J|^{-1}))$. This self-trapping dynamics is shown also in panel (b) for $t>10 |J|^{-1}$.
Another indicator of the change of the type of the dynamics is the change in the oscillation frequency, which (at least for small oscillations around the fixed points) can be calculated by the linear stability analysis of the fixed points, as summarized in Appendix \ref{sec:stabanal}. In Fig. \ref{fig:freq} we plot the oscillation frequency as a function of $I_0$. As we increase $I_0$ at the left hand side of the resonance, the Josephson oscillation frequency $\omega_J$ starts to grow first, since $J$ increases, and then at the right hand side where $J<0$ it decreases again, since $|J|$ decreases. Then at some point, when $U$ becomes bigger than $2|J|$ the fixed point for the Josephson oscillation becomes unstable and instead the self-trapping frequency $\omega_{\text{ST}}$ appears.
\begin{figure}[b!]
\centering
\includegraphics[width=\columnwidth]{lambdas}
\caption{(Color online) The oscillation frequencies as function of the strength of the potential dip. The given fixed point becomes unstable, when the curve goes below zero.}
\label{fig:freq}
\end{figure}
\section{Summary}
\label{sec:sum}
In this paper we have considered the effect of an additional central well added to the center of the symmetric double well barrier. We have shown that by suddenly opening up this narrow central well the tunneling amplitude of the bosonic Josephson junction can be ``quenched'' to almost arbitrary values. Therefore in experiments one can have an additional tunable parameter on the double well system and change the dynamics in-situ from plasma oscillations to the a.c Josephson dynamics or even to self-trapping without modifying the scattering properties.
\section*{Acknowledgements}
GSZ acknowledges support from the Hungarian National Office for Research and Technology under the contract ERC\_HU\_09 OPTOMECH, the Hungarian Academy of Sciences (Lend\"ulet Program, LP2011-016), the Hungarian Scientific Research Fund (grant no. PD104652) and the J\'anos Bolyai Scholarship. GM and LS acknowledge financial support from Università di Padova (Progetto di Ateneo grant No. CPDA 118083), Cariparo Foundation (Eccellenza grant 2011/2012), and MIUR (PRIN grant No. 2010LLKJBX).
|
\section{Introduction}
Let $\rho\in C^\infty({\mathbb {R}}^d\setminus \{0\})$ be a homogeneous distance function of degree $\beta>0$, i.e. $\rho$ satisfies
$\rho(t^{1/\beta}\xi)=t\rho(\xi)$ for all $t>0$ and $\rho(\xi)>0$ for $\xi\neq 0$.
For Schwartz functions $f$ on ${\mathbb {R}}^d$
($d\ge 2$ throughout this paper)
define the Riesz means $S^\lambda} \def\La{\Lambda_t f$ of the Fourier integral by
$$S^\lambda} \def\La{\Lambda_tf(x)=\frac{1}{(2\pi)^d}\int_{\rho(\xi)\le t}\Big(1-\frac{\rho(\xi)}{t}\Big)^\lambda} \def\La{\Lambda
\widehat f(\xi)e^{i\inn x\xi}\,d\xi.$$
In order to consider almost everywhere convergence on $L^p$ one needs to be able to define
$S^\lambda} \def\La{\Lambda_t f$ as a measurable function, for all $f\in L^p({\mathbb {R}}^d)$.
A necessary condition for $S^\lambda} \def\La{\Lambda_t$ to extend as a continuous
operator from $L^p$ to the space ${\mathcal {S}}'$ of tempered distributions
is that the convolution kernel belongs to $L^{p'}$ ({\it cf}.
\cite{caso} for a similar comment, and \S\ref{BRdef} below). This is
the case if and only if $\lambda} \def\La{\Lambda>\lambda} \def\La{\Lambda(p):=d(\frac 12-\frac1p)-\frac12$.
In view of the compact support of the multiplier the
distribution $S^\lambda} \def\La{\Lambda_t f$ is then a bounded $C^\infty$ function; moreover the maximal function $\sup_{t>0} |S^\lambda} \def\La{\Lambda_t f(x)|$ is well defined as a measurable function.
Carbery, Rubio de Francia and Vega \cite{crv} studied the pointwise behavior of $S^\lambda} \def\La{\Lambda_t$ and showed that, for $\rho(\xi)=|\xi|$ and $\lambda} \def\La{\Lambda>\max\{\lambda} \def\La{\Lambda(p),0\}$,
the means $S^\lambda} \def\La{\Lambda_t f(x)$ converge almost everywhere to $f(x)$.
The method in \cite{crv}
is based on the trace theorem for Sobolev functions and
applies to the general situation considered here, see also
Sato \cite{sato}. For an extension involving nonisotropic distance functions
see Cladek \cite{cladek}.
In this paper we investigate what happens at the critical index
$\lambda} \def\La{\Lambda=\lambda} \def\La{\Lambda(p)=d(\frac12-\frac1p)-\frac12$ when $\lambda} \def\La{\Lambda(p)>0$, {\it i.e.}
when $p>\frac{2d}{d-1}$. It is natural either to consider $S^\lambda} \def\La{\Lambda_t$
on smaller Lorentz spaces, or to slightly regularize the multiplier
to produce an operator well defined on $L^p({\mathbb {R}}^d)$. Let
\begin{equation}\label{hlaga}h_{\lambda} \def\La{\Lambda,\gamma}(r)= \frac{(1-r)^\lambda} \def\La{\Lambda_+}
{(1+\log(\tfrac{1}{1-r}))^\gamma}\,.\end{equation} Define the generalized Bochner
Riesz means $S^{\lambda} \def\La{\Lambda,\gamma}_tf$ by
$$\widehat {S^{\lambda} \def\La{\Lambda,\gamma}_t\! f}(\xi)= h_{\lambda} \def\La{\Lambda,\gamma}(\tfrac{\rho(\xi)}{t}) \widehat f(\xi).$$
For the critical $\lambda} \def\La{\Lambda=\lambda} \def\La{\Lambda(p)$
these operators
extend to bounded operators from the Lorentz space $L^{p,q}$ to ${\mathcal {S}}'$ if
and only if $\gamma>1-1/q$ when $q>1$ and $\gamma\ge 0$ when $q=1$;
{\it cf}. \S\ref{BRdef}.
In the thesis \cite{annoni} Annoni showed, for $\rho(\xi)=|\xi|$, that
$S^{\lambda} \def\La{\Lambda(p),\gamma}_t\! f(x)\to f(x)$ a.e. for all $f\in L^p$, $p>\frac{2d}{d-1}$
under the condition $\gamma> 3/2-1/p$. This left open the range
$1-1/p<\gamma\le 3/2-1/p$. Almost everywhere convergence in this range is implied
by the case $q=p$ of the following theorem.
\begin{theorem}\label{laga}
Let $\frac{2d}{d-1}<p<\infty$,
$\lambda} \def\La{\Lambda(p)=d(\frac 12-\frac 1p)-\frac12$.
Let $\gamma>1-\frac 1q$, $1<q\le \infty$, or $\gamma\ge 0$ and $q=1$. Then for
$f\in L^{p,q}$ we have
$$\lim_{t\to\infty} S^{\lambda} \def\La{\Lambda(p),\gamma}_t f(x)=f(x) \text{ a.e. }$$
\end{theorem}
\noindent Note that for $q=1$ this covers an endpoint result for the
Riesz means.
The usual approach to pointwise convergence
results for functions in $L^{p,q}({\mathbb {R}}^d)$ is to prove
an $L^{p,q}\to L^{p,\infty}$ bound for the associated maximal function
$S^{\lambda} \def\La{\Lambda,\gamma}_*\!f= \sup_{t>0}|S^{\lambda} \def\La{\Lambda,\gamma}_t\!f|$. By Stein's theorem
\cite{steinlimits}
such an estimate is necessary for $p\le 2$, see Tao \cite{tao} for
what is currently known for the maximal Bochner-Riesz operator in that range.
For $p>2$ complete $L^p$-boundedness results for the maximal Bochner-Riesz operator ($\lambda} \def\La{\Lambda>\lambda} \def\La{\Lambda(p)$) are known in two dimensions (\cite{carbery})
and partial results have been proved in higher dimensions (\cite{christ}, \cite{seegerthesis}, \cite{lrs-proc}, \cite{leese14}).
For the endpoint $\lambda} \def\La{\Lambda=\lambda} \def\La{\Lambda(p)$ sharp results were recently observed in
\cite{lrs}, for the case $\rho(\xi)=|\xi|$,
in the restricted range
$\frac{2d+2}{d-1}<p<\infty$. For these parameters
$S^{\lambda} \def\La{\Lambda(p),0}_*$
maps $L^{p,1} $ to $L^p$ and
$S^{\lambda} \def\La{\Lambda(p),\gamma}_*$ maps $L^{p,q}$ to $L^p$ if
$q\le p$ and $\gamma>1-1/q$. We note that
the sharp $L^{p,1}\to L^p$ endpoint bounds for the maximal operator are not known
in the range $\frac{2d}{d-1}< p\le \frac{2(d+1)}{d-1}$, even in two dimensions. Still less is known for general $\rho$.
Thus we opt for a variant of
the approach by Carbery, Rubio de Francia and Vega \cite{crv}, who used
weighted $L^2(|x|^{-a}dx)$ spaces.
Annoni \cite{annoni} followed this approach and introduced
logarithmic modifications of the weight function, working with
$|x|^{-a} (\log(2+ |x|))^{-\mu}$ for suitable $\mu>0$.
We prefer to preserve the
homogeneity of the weight and use the observation that for $p>2$
the space $L^{p,2}$ is embedded in $L^2(|x|^{-d(1-2/p)}dx)$. We are able to sharpen the analysis in \cite{crv} to prove boundedness in this space
for maximal operators defined by
$\sup_{t>0}|{\mathcal {F}}^{-1}[h(\tfrac{\rho(\cdot)}{t})\widehat f\,]|$ where $h$ is
supported in $[\tfrac 12,2]$ and belongs to the $L^2$-Sobolev space
${\mathscr {L}}^2_\alpha\equiv B^2_{\alpha,2}$ with $\alpha=d(1/2-1/p)$.
This will be a special case of Theorem \ref{herzest} below
and can also be deduced from the square-function estimate in
Theorem \ref{sq}.
In particular for the range $\gamma>1/2$ the maximal operator
$S^{\lambda} \def\La{\Lambda(p),\gamma}_*$ is bounded on
$L^2(|x|^{-d(1-2/p)}dx)$; this implies the
$q=2$ case
of Theorem \ref{laga}.
In order to get a complete result for all $q$
(in particular $q=p$)
it is convenient to work with
the homogeneous Herz spaces
$\hzr{\gamma}{r}$, in
the special case $r=2$. For fixed $r$ these are real interpolation spaces of the $L^r(|x|^{r\gamma}dx)$ spaces, see Gilbert \cite{gilbert}.
The definition (following the terminology in \cite{baernstein-sawyer})
is as follows. Let, for $l\in {\mathbb {Z}}$,
${\mathfrak {A}}_l:=\{x:2^l\le |x|<2^{l+1}\}$. Then $\hzr{\gamma}{q}$ is
the space of all functions which are $r$-integrable on compact subsets of ${\mathbb {R}}^d\setminus\{0\}$ and for which
\begin{equation}\label{herzdef}
\|f\|_{\hzr{\gamma}{q}}=
\Big(\sum_{l\in {\mathbb {Z}}}2^{l\gamma q} \Big[ \int_{{\mathfrak {A}}_l} |f(x)|^r\, dx\Big]^{q/r}\Big)^{1/q}
\end{equation}
is finite. It is easy to see that for $\gamma>-d/r$ every function in
$\hzr{\gamma}{q}$ defines a unique tempered distribution on ${\mathcal {S}}'({\mathbb {R}}^d)$.
In what follows we use the notation $\lesssim$ for an inequality which involves an implicit constant.
\begin{theorem} \label{herzest}
Let $\frac12<\alpha<\frac d2$, $1\le q\le \infty$ and
$s=\frac{q}{q-1}$. Let $h\in B^2_{\alpha,s}({\mathbb {R}})$ be supported in
$(\frac 12, 2)$. Then the following hold.
(i) For $2\le q\le \infty$,
$$
\big\| \sup_{t>0} \big|{\mathcal {F}}^{-1}[h(\tfrac{\rho(\cdot)}{t}) \widehat
f\,]
\big\|_{\hz{-\alpha}{q}}
\lesssim \|h\|_{B^2_{\alpha,s}}
\|f\|_{\hz{-\alpha}{q}}\,.
$$
(ii) For $1\le q\le2$,
$$
\big\| \sup_{t>0} \big|{\mathcal {F}}^{-1}[h(\tfrac{\rho(\cdot)}{t}) \widehat
f\,]
\big\|_{\hz{-\alpha}{2}}
\lesssim \|h\|_{B^2_{\alpha,s}}
\|f\|_{\hz{-\alpha}{q}}\,.
$$
\end{theorem}
For $p>2$ and $\alpha=d(1/2-1/p)$ there is the embedding
$L^{p,q}\subset \hz{-\alpha}{q}$, ({\it cf}. Lemma \ref{Loremb} below).
By a standard approximation argument we will show
\begin{corollary}\label{herzcor}
Let $p>\frac{2d}{d-1}$, $1\le q\le\infty$, $s=\frac{q}{q-1}$,
and let $h\in B^2_{d(\frac 12-\frac1p),s}$
be supported in a compact subinterval of $(0,\infty)$. Then
for $f\in L^{p,q}$
$$\lim_{t\to\infty} |{\mathcal {F}}^{-1}[h(\rho/t)\widehat f\,](x)|=0 \text{ a.e.}$$
\end{corollary}
If $\chi\in C^\infty_0({\mathbb {R}})$ is compactly supported away from the origin
then $\chi h_{\lambda} \def\La{\Lambda,\gamma}\in B^2_{\alpha,s}$ for $\alpha=\lambda} \def\La{\Lambda+1/2$ and $\gamma>1/s$, and $\gamma\ge 0$ if $s=\infty$, see {\it e.g. }
\cite{taibleson}. Thus one
can deduce Theorem \ref{laga} from Corollary \ref{herzcor}.
The proof of Theorem \ref{herzest} for $q=2$ also
gives a characterization of
boundedness convolution operators with quasiradial multipliers on
weighted $L^2$-spaces with power weights (i.e. on the spaces
$\dot {\mathscr {K}}^{2}_{\pm \alpha,2}$).
The result is in the spirit of (but in some sense complementary to)
the results by
Muckenhoupt, Wheeden and Young \cite{mwy}.
In what follows let $\varphi\in C^\infty_0(\tfrac 12, 2)$ be a nontrivial
bump function.
\begin{theorem} \label{multch}
Let $\rho$ be as above and let $1/2<\alpha<d/2$. Define the operator $T$
by $\widehat {Tf}(\xi)= m(\rho(\xi))\widehat f(\xi)$. Then the following statements are equivalent:
(i) $T$ is bounded on $L^2(|x|^{-2\alpha} dx)$.
(ii) $T$ is bounded on $L^2(|x|^{2\alpha} dx)$.
(iii) $\sup_{t>0}\|\varphi m(t\cdot)\|_{L^2_\alpha({\mathbb {R}})} <\infty$.
\end{theorem}
Finally we prove a sharp bound for Stein's square-function
associated with the Riesz means,
$$G_\alpha f(x)= \Big(\int_0^\infty
\big| S^{\alpha-1}_t f(x)-S^\alpha_t f(x)\big|^2\frac{dt}{t}\Big)^{1/2}\,$$
which can also be used to prove Theorem \ref{multch}.
\begin{theorem}\label{sq}
Let $\frac 12<\alpha<\frac d2$.
Then for $f\in L^2(|x|^{-2\alpha}dx)$,
$$\int\big| G_\alpha f(x)\big|^2 \frac{dx}{|x|^{2\alpha}}\lesssim
\int \big|f(x)\big|^2 \frac{dx}{|x|^{2\alpha}}\,.$$
\end{theorem}
\medskip
\subsection*{\it This paper.} In \S\ref{BRdef} we discuss the definition
of our multiplier transformations and associated maximal functions
on Lebesgue and Lorentz spaces, and then prove a convenient
characterization of quasiradial multipliers to be Fourier transforms
of functions in these spaces (see Theorem \ref{FLuprop}). In the
preliminary section \S\ref{prel} we consider embeddings for Lorentz
and Herz spaces which are needed to deduce Theorem \ref{laga} from
Corollary \ref{herzcor}. We also review the Fourier restriction
estimate in weighted $L^2$ spaces, use it to prove a basic maximal
function estimate, and discuss other basic preliminaries. The main
section \S\ref{herzestsect} contains the proof of Theorem
\ref{herzest}. The proofs of Theorem \ref{multch} and Corollary
\ref{herzcor} are given in \S\ref{concl}, and Theorem \ref{sq} is
proved in \S\ref{sqsect}.
\subsection*{\it Acknowledgement} The authors would like to thank Jong-Guk Bak for conversations
at the early stages of this work.
\section{Definition of convolution and maximal operators
on $L^{p,q}$}\label{BRdef}
For $f\in L^p$ the expression $S^{\lambda} \def\La{\Lambda, \gamma}_t f$ is not necessarily
defined for all $f\in L^{p,q}$, even in the sense of tempered distributions, since
the Fourier transform of an $L^p$ function does not have to be a function.
\begin{lemma} \label{TKlemma} Suppose $1\le p,q<\infty$. Let $K$ be a tempered distribution in ${\mathbb {R}}^d$ such that $\widehat K$ has compact support. For $f\in {\mathcal {S}}({\mathbb {R}}^d)$ define $T_K=f*K$. Then
(i) $T_K$ extends to a continuous linear operator from $L^{p,q}$ to ${\mathcal {S}}'$ if and only if $K\in L^{p',q'}$.
(ii) If $K\in L^{p',q'}$ then for $f\in L^{p,q}$ the maximal function
$$\sup_{t>0}|t^{d}K(t\cdot)* f(x)|$$ is Borel measurable.
\end{lemma}
\begin{proof}
Assume that $T_K: L^{p,q}\to {\mathcal {S}}'$ is continuous.
Then for every $g\in {\mathcal {S}}$ there
is a constant $C(g)$ such that $|\inn{T_K f}{g}|\le
C(g)\|f\|_{L^{p,q}}$. Observe that if $g\in {\mathcal {S}}$ such that $\widehat
g(\xi)=1$ for $\xi\in {\hbox{\roman supp}}(\widehat K)$ then $\inn{K*f}{g}=
\inn{f}{K(-\cdot)}$. Thus $K$ must lie in the dual space of
$L^{p,q}$, i.e. in $L^{p',q'}$. Conversely, if $K\in K^{p',q'}$
then for every $f\in L^{p,q}$ the convolution $x\mapsto K*f(x)$ is
well defined as a bounded continuous function. This shows $(i)$.
Since the supremum of continuous functions is Borel measurable we
get $(ii)$.
\end{proof}
\subsection*{Quasiradial functions in ${\mathcal {F}} L^{u,s}$}\label{quasiradial}
We consider $m=h\circ\rho$ where $\rho$ is a homogeneous distance
function of degree $\beta$, and $\rho\in C^\infty({\mathbb {R}}^d\setminus
\{0\})$. It is useful to express the condition $(ii)$ in Lemma
\ref{TKlemma} in terms of the one-dimensional Fourier transform of
$h_\beta(t)= h(t^\beta)$. The following theorem sharpens a result
in \cite{se-toh} for the class of distance functions considered
here; the radial analogue, for $\rho(\xi)=|\xi|$, is already in
\cite{gs}. Note that here we make no curvature assumption on the $\rho$-unit sphere
$$\Sigma_\rho=\{\xi:\rho(\xi)=1\}.$$ The following result may be interesting in its own right but we
shall use it only to demonstrate sharpness of our
results; it is not needed in the proofs of Theorems
\ref{herzest}-\ref{sq}.
\begin{theorem} \label{FLuprop}Let $h$ be supported in a compact subinterval $J$ of $(0,\infty)$ and let $${\mathcal {K}}_{\beta}(r) = \frac{1}{2\pi}\int h(t^\beta) e^{i rt} dt\,.$$
Let $1<u<2$, $1\le s\le \infty$ and $\mu_d$ be the measure on ${\mathbb {R}}$ given by
$$d\mu_d(r)= (1+|r|)^{d-1} dr\,.$$ Then
\begin{equation}\label{flqeq}\big\|{\mathcal {F}}^{-1} [h\!\circ\!\rho] \big\|_{L^{u,s}({\mathbb {R}}^d)} \approx
\Big\| \frac{ {\mathcal {K}}_{\beta } }{(1+|\cdot|)^{\frac{d-1}{2}}} \Big\|_{L^{u,s}({\mathbb {R}}, \mu_d)}.\end{equation}
\end{theorem}
For $u=s$ this equivalence becomes
$$\big\|{\mathcal {F}}^{-1} [h\!\circ\!\rho] \big\|_{L^{u}({\mathbb {R}}^d)} \approx
\Big(\int |{\mathcal {K}}_{\beta}(r)|^u (1+|r|)^{(d-1)(1-\frac u2)}
dr\Big)^{1/u}\,.$$
In the proof of Theorem \ref{FLuprop} we shall use an elementary convolution inequality in weighted spaces.
\begin{lemma} \label{weightineq}For $a\in {\mathbb {R}}$ let $M_a$ be the multiplication operator $M_a g(r)=g(r)(1+|r|)^a$.
Let $\varsigma$ be a measurable function such that for all $N$ $$|\varsigma(r)|\le C(N)(1+|r|)^{-N}.$$ Then for all $a\in {\mathbb {R}}$, $1\le u\le\infty$
$$\big\| M_{-a} [\varsigma * (M_a g)]\big\|_{L^{u,s}(\mu_d)} \lesssim \|g\|_{L^{u,s}(\mu_d)}\,.
$$
\end{lemma}
\begin{proof} The straightforward proof is left to the reader. Real interpolation reduces this to the case
$u=s$ for which one can consult Lemma 2.2 in \cite{gs}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{FLuprop}]
Since $\rho^{1/\beta}$ is homogeneous of degree $1$ the statement follows immediately from the special case $\beta=1$, which we will henceforth assume. We shall write $\kappa$ for ${\mathcal {K}}_1$, so that $\kappa={\mathcal {F}}^{-1}_{\mathbb {R}}[h]$.
\medskip
\noindent{\it Proof of $\lesssim$ in \eqref{flqeq}.}
Let $\chi_1$ in $C^\infty$, compactly supported in $(0,\infty)$ such that $\chi_1(\rho)=1$ on the support of $h$. Since $\kappa=\widehat h$ it suffices to show
$$
\big\|{\mathcal {F}}^{-1}\big[ \chi_1\!\circ\!\rho\, \,\widehat\kappa\!\circ\!\rho \big]
\big\|_{L^{u,s}({\mathbb {R}}^d)} \lesssim
\big\|(1+|\cdot|)^{-\frac{d-1}{2}}\kappa\big\|_{L^{u,s}(\mu_d)}
$$
which follows from \begin{equation}\label{suffest} \Big\|{\mathcal {F}}^{-1} \big[
\chi_1\!\circ\!\rho \int g(r)(1+|r|)^{\frac{d-1}{2}} e^{-ir\rho} dr
\big] \Big\|_{L^{u,s}({\mathbb {R}}^d)} \lesssim \big\|g\big\|_{L^{u,s}(\mu_d)}.
\end{equation}
We show the corresponding inequalities with $L^{u,s}$ replaced
by $L^1$ and by $L^2$.
For the $L^1$ inequality we use that
$$\big\|{\mathcal {F}}^{-1} \big[ \chi_1\!\circ\!\rho\, \, e^{-ir\rho} \big] \big\|_{L^1({\mathbb {R}}^d)}
\lesssim (1+|r|)^{\frac{d-1}{2} }$$
which is a rescaled inequality from \cite{sesost}. Here it is crucial that $\rho$
is homogeneous of degree one. By Minkowski's inequality the displayed estimate yields
\begin{equation}\label{L1suff}
\Big\|{\mathcal {F}}^{-1} \big[ \chi_1\!\circ\!\rho \int g(r)(1+|r|)^{\frac{d-1}{2}} e^{-ir\rho} dr \big]
\Big\|_{L^1({\mathbb {R}}^d)} \lesssim \int |g(r)| (1+|r|)^{d-1} dr
\end{equation}
For the $L^2$ inequality we use polar coordinates and Plancherel's theorem in ${\mathbb {R}}^d$ and ${\mathbb {R}}$,
to get
\begin{align}
\notag
&\Big\|{\mathcal {F}}^{-1} \big[ \chi_1\!\circ\!\rho \int g(r)(1+|r|)^{\frac{d-1}{2}} e^{-ir\rho} dr \big]
\Big\|_{L^2({\mathbb {R}}^d)}
\\ \notag &\lesssim
\Big\|\chi_1\!\circ\!\rho(\cdot) \int g(r)(1+|r|)^{\frac{d-1}{2}} e^{-ir\rho(\cdot)} dr
\Big\|_{L^2({\mathbb {R}}^d)}
\\ \notag &\lesssim
\Big\|\chi_1 \int g(r)(1+|r|)^{\frac{d-1}{2}} e^{-ir(\cdot)} dr
\Big\|_{L^2({\mathbb {R}})}
\\ \label{L2suff}
&\lesssim \Big(\int |g(r)|^2 (1+|r|)^{d-1} dr \Big)^{1/2}\,.
\end{align}
The asserted inequality \eqref{suffest} follows from \eqref{L1suff} and \eqref{L2suff} by real interpolation.
\medskip
\noindent{\it Proof of $\gtrsim$ in \eqref{flqeq}.}
We shall work with smooth $h$ which allows
to assume that the $L^{u,s}(\mu_d)$ norm of $(1+|r|)^{-\frac{d-1}{2}}\kappa(r)$ is a priori finite.
With this assumption we have to prove the inequality
\begin{equation}\label{necpart}
\big\| (1+|\cdot|)^{-(d-1)/2} \kappa\|_{L^{u,s}(\mu_d)}
\lesssim \|{\mathcal {F}}^{-1}[ h\circ\rho]\|_{L^{u,s}({\mathbb {R}}^d)} \,.
\end{equation}
If \eqref{necpart} holds for all smooth $h$ supported in $J$ then the general case can be derived by an approximation argument.
Pick $\xi_0\in \Sigma_\rho$ so that $|\xi|$ has a maximum at
$\xi_0$. Then the Gauss map is injective in a small neighborhood $U$ on the surface $\Sigma_\rho$ and the curvature is bounded below on $U$. Let $\gamma$ be homogeneous of degree zero, $\gamma(\xi_0)\neq 0$ and supported on the closure of the cone generated by $U$. Clearly
$$\|{\mathcal {F}}^{-1} [\gamma\, h\!\circ\!\rho
]\|_{L^{u,s}}\lesssim
\|{\mathcal {F}}^{-1} [h\!\circ\!\rho]\|_{L^{u,s}}.$$
Now use $\rho$ polar coordinates to write
\begin{equation}\label{polar}
{\mathcal {F}}^{-1} [\gamma\, h\!\circ\!\rho ](x)=
(2\pi)^{-d} \int_0^\infty h(\rho)\rho^{d-1}\int_{\Sigma_\rho}\gamma(\xi') e^{i\rho\inn{\xi'}{x}}
\frac{d\sigma(\xi')}{|\nabla\rho(\xi')|} \,d\rho\,.
\end{equation}
Let $n(\xi_0)$ the outer normal at $\xi_0$, let $\Gamma=\{x\in {\mathbb {R}}^d: \big|\frac{x}{|x|}-n(\xi_0)\big|
\le \varepsilon\}$, with $\varepsilon$ small and let, for large $R\gg1 $, $\Gamma_R=\{x\in\Gamma:|x|\ge R\}$.
By choosing $\varepsilon$ small enough we may assume that for each $x\in \Gamma$ there is a unique $\xi=\Xi(x)\in \Sigma_\rho$, so that $\gamma(\Xi(x))\neq 0$ and so that $x$ is normal to $\Sigma_\rho$ at $\Xi(x)$. Clearly $x\mapsto\Xi(x)$ is homogeneous of degree zero on $\Gamma$.
By the method of stationary phase we have for $x\in \Gamma_R$
\begin{equation} \label {gahdec}{\mathcal {F}}^{-1} [\gamma h(\rho(\cdot))](x)= I_0(x)+\sum_{j=1}^N II_j(x)+III(x)\end{equation}
where
\begin{align*}
I_0(x)&=c\int_0^\infty h(\rho)\rho^{\frac{d-1}{2}} e^{i\rho\inn{\Xi(x)}{x}} d\rho
\frac{\gamma(\Xi(x)) |\nabla\rho(\Xi(x))|^{-1}}
{(\inn{\Xi(x)}{x})^{\frac{d-1}{2}}
|K(\Xi(x))|^{1/2}},
\end{align*}
$|c|=(2\pi)^{-d}$ and $K(\Xi(x))$ is the curvature of $\Sigma_\rho$ at $\Xi(x)$.
There are similar formulas for the higher order terms $II_j(x)$ with
the main term $(\rho\inn{\Xi(x)}{x})^{-\frac{d-1}{2}}$ replaced by
$(\rho\inn{\Xi(x)}{x})^{-\frac{d-1}{2}-j}$. Finally
$$| III (x) |\lesssim_N \|h\|_1|x|^{-N}\,, \quad x\in \Gamma_R\,.$$
Let $h_j(\rho)=h(\rho)\rho^{\frac{d-1}{2}-j}$ and let $\kappa_j ={\mathcal {F}}^{-1}_{\mathbb {R}}[h_j]$, then,
for some $C_1\ge 1$,
$$C_1^{-1}| I_0(x) | \le \frac{|\kappa( \inn{\Xi(x)}{x}) |}
{|\inn{{\Xi(x)}{x}}|^{\frac{d-1}{2}} } \,\le C_1| I_0(x)|\,,
\quad x\in \Gamma_R.$$
We also have, for some $C_0\ge 1$,
$$ C_0^{-1}|x|\le \inn{\Xi(x)}{x}\le C_0 |x|, \quad x\in \Gamma,$$ which is a consequence of Euler's homogeneity relation $\rho(\xi)= \inn{\xi}{\nabla\rho(\xi)}$ and the positivity assumption on $\rho$.
Let \begin{align*}
E_R(\theta,\alpha)&= \{r: |r|\ge R, | I_0(r\theta) |> \alpha\}\,,
\\
E_{R}^*(\beta)&=\{r: |r|\ge C_0R: \, r^{-\frac{d-1}{2}}|\kappa_0(r)|>C_1\beta\}\,.
\end{align*}
Then
\begin{align*}
\| I_0\|_{L^{u,s}(\Gamma_R)}
&\gtrsim\Big(\int_0^\infty \big[\alpha\,({\text{\rm meas}}\{x\in \Gamma_R:| I_0(x)|>\alpha\})^{1/u}\big]^s \frac{d\alpha}\alpha\Big)^{1/s}
\\
&\gtrsim \Big(\int_0^\infty \Big[\alpha \Big(\int_{S^{d-1}\cap\Gamma} \int_{E_R(\theta,\alpha)} r^{d-1} dr \,d\theta\Big)^{1/u}\Big]^s \frac{d\alpha}\alpha\Big)^{1/s}
\\
&\gtrsim \Big(\int_0^\infty \Big[\alpha \Big(\int_{E_{R}^*(2^{d+1}\alpha)} r^{d-1} dr \Big)^{1/u}\Big]^s
\frac{d\alpha}\alpha\Big)^{1/s},
\end{align*}
which gives
\begin{equation} \label{I0below}\big\| (1+|\cdot|)^{-\frac{d-1}{2}}\kappa_0\big\|_{L^{u,s}([C_0 R,\infty],\mu_d)}
\lesssim \| I_0\|_{L^{u,s}(\Gamma_R)} \,.
\end{equation}
A variant of this argument also yields the upper bound
\begin{equation} \notag
\| I_0\|_{L^{u,s}(\Gamma_R)} \lesssim
\big\| (1+|\cdot|)^{-\frac{d-1}{2}}\kappa_0\big\|_{L^{u,s}(\mu_d)}\,,
\end{equation}
and similarly, taking into account the additional decay in the terms $II_j$,
\begin{align}\notag
\| II_j\|_{L^{u,s}(\Gamma_R)} &\lesssim R^{-1}
\big\| (1+|\cdot|)^{-\frac{d-1}{2}}\kappa_j\big\|_{L^{u,s}(\mu_d)}
\\ \label {upperboundsII}
&\lesssim R^{-1}
\big\| (1+|\cdot|)^{-\frac{d-1}{2}}\kappa_0\big\|_{L^{u,s}(\mu_d)}\,.
\end{align}
Here, for the second inequality, we have used Lemma \ref{weightineq} and the fact that $\kappa_j=\kappa*\varsigma_j$ for some Schwartz function $\varsigma_j$.
We also have the trivial inequalities which use the support properties of $h$
$$\| III\|_{L^{u,s}(\Gamma_R)} \lesssim R^{-1} \|h\|_1$$
and
$
\|h\|_1 \lesssim \|h\!\circ\!\rho\|_{L^{u',s}({\mathbb {R}}^d)}
\lesssim
\|{\mathcal {F}}^{-1}[h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)}\,,
$
by the Hausdorff-Young inequality. Thus
\begin{equation}\label{three}
\| III\|_{L^{u,s}(\Gamma_R)} \lesssim R^{-1} \|{\mathcal {F}}^{-1}[h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)}.
\end{equation}
This argument also gives
\begin{equation} \label{I0belowtriv}\big\| (1+|\cdot|)^{-\frac{d-1}{2}}\kappa_0\big\|_{L^{u,s}([0,C_0],\mu_d)}
\lesssim R^d \| {\mathcal {F}}^{-1}[h\!\circ\!\rho]\|_{L^{u,s}}
\end{equation}
since the left hand side is estimated by $ R^d \|\kappa_0\|_\infty \lesssim R^d \|h\|_1$.
We now combine the estimates to show \eqref{necpart}. By \eqref{I0below} and \eqref{I0belowtriv} we get
\begin{align*}
&\big\| (1+|\cdot|)^{-\frac{d-1}2} \kappa_{0} \|_{L^{u,s}(\mu_d)}
\lesssim \| I_0 \|_{L^{u,s} (\Gamma_R)} + R^d\|{\mathcal {F}}^{-1}[ h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)}.
\end{align*}
Since $\|{\mathcal {F}}^{-1}[\gamma\, h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)}
\lesssim \|{\mathcal {F}}^{-1}[ h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)} $, by \eqref{gahdec}, \eqref{three} and \eqref{upperboundsII}
\begin{align*}
&\| I_0 \|_{L^{u,s} (\Gamma_R)}\lesssim (1+R^d)\|{\mathcal {F}}^{-1}[h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)}+
\sum_{j=1}^N\| II_j \|_{L^{u,s} (\Gamma_R)}
\\ &\lesssim (1+R^d)\|{\mathcal {F}}^{-1}[h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)}+ R^{-1}
\big\| (1+|\cdot|)^{-\frac{d-1}{2}}\kappa_0\big\|_{L^{u,s}(\mu_d)}.
\end{align*}
Hence, combining these two estimates and choosing $R\gg1 $ sufficiently large we get
$$\big\| (1+|\cdot|)^{-\frac{d-1}2} \kappa_{0} \|_{L^{u,s}(\mu_d)}
\lesssim \|{\mathcal {F}}^{-1}[ h\!\circ\!\rho]\|_{L^{u,s}({\mathbb {R}}^d)} \,.$$
Finally observe that $\kappa= \kappa_0*\varsigma_0$ for some Schwartz function $\varsigma_0$ and thus by Lemma \ref{weightineq}
$$\big\| (1+|\cdot|)^{-\frac{d-1}2} \kappa \|_{L^{u,s}(\mu_d)}
\lesssim\big\| (1+|\cdot|)^{-\frac{d-1}2} \kappa_{0} \|_{L^{u,s}(\mu_d)}\,.$$
The desired estimate \eqref{necpart} follows.
\end{proof}
\subsection*{Quasiradial functions in Besov spaces}
If $1< u< 2$ we have the embedding $\hz{\alpha}{s}
\hookrightarrow L^{u,s}$
for $\alpha= d(1/u-1/2)$. This holds for $u=1$, $s=1$ by using the Cauchy-Schwarz inequality on each annulus ${\mathfrak {A}}_l$, and it trivially holds for $u=s=2$.
We use interpolation (\cite{gilbert}) together with the reiteration theorem for real interpolation to deduce that the embedding holds for $1<u<2$ and all $s>0$.
Next observe that the homogeneous Besov space $\dot B^2_{\alpha,s}$ is the image of $\hz{\alpha}{s}$ under the Fourier transform. If $h$ is supported in a compact subinterval $J$ of $(0,\infty)$ then for $\alpha>0$,
$$ \|h\!\circ\!\rho\|_{B^2_{\alpha,q}({\mathbb {R}}^d)}
\lesssim \|h\|_{B^2_{\alpha,q}({\mathbb {R}})} \lesssim \|h\|_{\dot B^2_{\alpha,q}({\mathbb {R}})};
$$
here the implicit constants depend on $J$.
Hence we see that if $p>2$ and $h\in B^2_{d(1/2-1/p),q'}$ is supported in $J$
then ${\mathcal {F}}^{-1}[h(\rho)\widehat f\,]$
and the associated maximal function $\sup_{t>0}
|{\mathcal {F}}^{-1}[h(\rho/t)\widehat f\,]$ are well defined for $f\in
L^{p,q}$.
\subsection*{Multipliers of Bochner-Riesz type}
Now let $\{\eta_j\}$ be a family of $C^\infty$-functions supported
on $(1/4,1/2)$ so that the $C^{10 d}$ norms are uniformly bounded.
Define, for sequences ${\mathfrak {a}}=\{a_j\}_{j=2}^\infty$
\begin{equation}\label{ha}h[{\mathfrak {a}}](\tau) = \sum_{j=2}^\infty a_j 2^{-j\lambda} \def\La{\Lambda}
\eta_j(2^j(1-\tau)). \end{equation} Then it is easy to see that for $\lambda} \def\La{\Lambda>-1/2$
\begin{equation} \label{sequences}\|h[{\mathfrak {a}}]\|_{B^2_{\lambda} \def\La{\Lambda+\frac 12, s}} \lesssim
\|{\mathfrak {a}}\|_{\ell_s}. \end{equation}
Now consider $h_{\lambda} \def\La{\Lambda,\gamma}$ in \eqref{hlaga} and let
$\chi\in C^\infty$ be supported in $(9/10, 11/10)$ and equal to $1$ near $1$.
It is well known that $[(1-\chi) h_{\lambda} \def\La{\Lambda,\gamma}]\circ\rho$ is the
Fourier transform
of an $L^1$ function and that the associated maximal operator is of
weak type $(1,1)$.
The function $\chi h_{\lambda} \def\La{\Lambda,\gamma}$ can be written as
in \eqref{ha} for $a_j= j^{-\gamma}$. Thus
$\chi h_{\lambda} \def\La{\Lambda,\gamma}
\in B^2_{\lambda} \def\La{\Lambda+1/2,s}$ for $\gamma s>1$.
Hence $S^{\lambda} \def\La{\Lambda,\gamma}_t f$ and
the associated maximal function are well
defined on $L^{p,q}$ if $\lambda} \def\La{\Lambda=d(1/2-1/p)-1/2$ and $\gamma>1-1/q$.
Finally to show lower bounds we consider the one dimensional inverse Fourier transform
$\kappa_{\lambda} \def\La{\Lambda,\gamma}$ of $\chi h_{\lambda} \def\La{\Lambda,\gamma} $, i.e.
\begin{align*}
2\pi \kappa_{\lambda} \def\La{\Lambda,\gamma}(r) &= \int \chi(v) (1-v)_+^\lambda} \def\La{\Lambda (\log \frac{1}{|1-v|})^{-\gamma} e^{ivr} \,dv
\\&= e^{ir} \int \chi_0(t) t_+ ^\lambda} \def\La{\Lambda (\log |t|^{-1})^{-\gamma} e^{-i tr}\,dt
\end{align*}
here $\chi_0$ is supported in $(-10^{-1},10^{-1})$.
By a standard asymptotic expansion (\cite{erd}, ch. 2.9.), we can estimate the expression that
we get by freezing the logarithmic term at $t=r^{-1}$, namely we have for $|r|\gg 1$
$$\Big|
\int \chi_0(t) t_+ ^\lambda} \def\La{\Lambda e^{-i tr}\,dt\, (\log r )^{-\gamma}
\Big| \,\gtrsim\,|r|^{-\lambda} \def\La{\Lambda-1} (\log |r|)^{-\gamma}.
$$
Straightforward estimation (using integration for the parts where $t\approx r^{-1}2^m$ with $m>0$)
shows
$$\Big|
\int \chi_0(t) t_+ ^\lambda} \def\La{\Lambda
\big[ (\log |t|^{-1})^{-\gamma}
- (\log r )^{-\gamma} \big]
e^{-i tr}\,dt
\Big| \lesssim |r|^{-\lambda} \def\La{\Lambda-1} (\log |r|)^{-\gamma-1}.
$$
The details are left to the reader. In view of the additional logarithmic gain in the last display we obtain for $|r|\gg1$
$$|\kappa_{\lambda} \def\La{\Lambda,\gamma}(r)|\gtrsim |r|^{-\lambda} \def\La{\Lambda-1} (\log |r|)^{-\gamma}\,.$$
From this it is easy to see that the condition
$$ \frac{\kappa_{\lambda} \def\La{\Lambda,\gamma}}{(1+|r|)^{\frac{d-1}{2}}} \in L^{p',q'}((1+|r|)^{d-1}dr )$$
implies that either
$\lambda} \def\La{\Lambda> d(1/2-1/p)-1/2$ or
$\lambda} \def\La{\Lambda=d(1/2-1/p)-1/2$ and $\gamma>1/q'$.
Thus Theorem \ref{weightineq} and Lemma \ref{TKlemma} show that the range of $\gamma$ in Theorem \ref{laga} is optimal.
\section{Preliminaries and basic estimates}\label{prel}
\subsection*{An embedding result}
Recall the notation ${\mathfrak {A}}_l=\{2^l\le|x|< 2^{l+1}\}$.
\begin{lemma}\label{Loremb} Let $0<a<d$, $1<r<\infty$ and let
$p=\frac{rd}{d-a}$.
Then, for $1<q<\infty$,
$$
\Big(\sum_{l\in \mathbb Z}\Big[\int_{{\mathfrak {A}}_l}
|f(x)|^r |x|^{-a}dx\Big]^{q/r}\Big)^{1/q} \lesssim
\|f\|_{L^{p,q}}\,,
$$
and
$$
\sup_l \Big(\int_{{\mathfrak {A}}_l}
|f(x)|^r |x|^{-a}dx\Big)^{1/r} \lesssim
\|f\|_{L^{p,\infty}}\,.
$$
\end{lemma}
\begin{proof}
Let $f$ be measurable so that $|f(x)|\le \chi_E(x)$ for some set $E$ of
finite Lebesgue measure.
Let $\varpi_d$ be the surface measure of the unit sphere in ${\mathbb {R}}^d$ and let
$B(E)$ be the ball centered at the origin, with radius
$R(E)= (d\varpi_d^{-1}|E|)^{1/d}$; then $E$ and $B(E)$ have the same Lebesgue measure and
$$
\int_E |x|^{-a} dx \le \int_{B(E)}|x|^{-a} dx = \frac{d}{d-a}
\Big(\frac{\varpi_d}{d}\Big)^{a/d}
|E|^{1-\frac ad}\,.$$
This implies the
inequality
$$
\Big(\int |f(x)|^r |x|^{-a}dx\Big)^{1/r}\lesssim
\|f\|_{L^{p,1}}\,,\quad \frac 1p= \frac 1r-\frac{a}{rd}\,.
$$
We apply this inequality
with two choices $(p_0,a_0)$ and $(p_1,a_1)$ satisfying
$a_i/r= d(1/r-1/p_i)$ for $i=0,1$ where $a_0>0$ is close to $0$ and $a_1<d$ is close to $d$.
Let $a_0<a<a_1$ and let $p=\tfrac{rd}{d-a}$. Then there is
$\vartheta\in (0,1)$ so that $(\tfrac ar,\tfrac 1p)=
(1-\vartheta) (\tfrac{a_0}r,\tfrac 1{p_0})+\vartheta (\tfrac{a_1}r,\tfrac 1{p_1})$.
Since $[L^{p_0,1}, L^{p_1,1}]_{\vartheta,q}= L^{p,q}$ and since $[L^r(|x|^{-a_0} dx),
L^r(|x|^{-a_1} dx)]_{\vartheta,q}= \hzr{-a/r}{q}$, by Gilbert's result
\cite{gilbert}, the assertion follows by interpolation.
\end{proof}
\subsection*{ Fourier restriction based on traces of
Sobolev spaces}
In what follows we let $\sigma$ be the surface measure on the $\rho$-unit sphere
$\Sigma_\rho$.
The following result is standard, but we include the proof for completeness.
\begin{lemma}\label{trace} For $1<b<d$,
$$\int_{\Sigma_\rho} |\widehat g(\xi)|^2 d\sigma(\xi)
\lesssim \int |g(x)|^2 |x|^b dx\,.
$$
\end{lemma}
\begin{proof} We split $g=g_0+g_1$ where $g_0(x)=0$ for $|x|>1$ and $g_1(x)=0$ for $|x|\le 1$.
Then
\begin{align*}
&\Big(\int_{\Sigma_\rho} |\widehat g_0(\xi)|^2 d\sigma(\xi)\Big)^{1/2}
\lesssim \|\widehat g_0
\|_\infty \lesssim \|g_0\|_1
\\&\lesssim \Big(\int|g_0(x)|^2 |x|^b \,dx\Big)^{1/2} \Big(\int_{|x|\le 1} |x|^{-b} dx\Big)^{1/2} \lesssim \Big(\int |g(x)|^2 |x|^b \,dx\Big)^{1/2}
\end{align*}
where we have used $b<d$.
Now the trace theorem says that for $b>1$
the restriction to $\Sigma_\rho$ of functions in the Sobolev space
${\mathscr {L}}^2_{b/2}({\mathbb {R}}^d)$
belongs to
${\mathscr {L}}^2_{(b-1)/2}(\Sigma_\rho)$, so it is in $L^2(\Sigma_\rho)$. We apply
the corresponding inequality to $\widehat g_1$ and combine it with Plancherel's theorem to see that
$$\Big(\int_{\Sigma_\rho} |\widehat g_1(\xi)|^2 d\sigma(\xi)\Big)^{1/2}
\lesssim \Big(\int |g_1(x)|^2 (1+|x|^2)^{b/2} dx\Big)^{1/2}.
$$ In view of the support of $g_1$ we may replace the weight
$(1+|x|^2)^{b/2}$ with $|x|^b$ and then $g_1$ with $g$. Finally, combine the estimates for
$\widehat g_0$ and $\widehat g_1$.
\end{proof}
\subsection*{Estimates for maximal functions}
We use Lemma \ref{trace} to bound a maximal operator associated with $h\circ \rho$.
\begin{proposition}\label{basicmult} Let $1<b<d$.
Let $\rho\in C^\infty({\mathbb {R}}^d\setminus \{0\}$, homogeneous of degree $\beta>0$ and positive
on ${\mathbb {R}}^d\setminus \{0\}$. Suppose that
$$ \Big( \int_0^\infty
|h(s)|^2s^{\frac b\beta-1}ds\Big)^{1/2} \le A.
$$
Then
$$\Big(\int \big|\sup_{1<t<2}|{\mathcal {F}}^{-1}[h(\rho(\cdot)/t)\widehat f\,]\big|^2\frac{dx}{|x|^b}\Big)^{1/2}
\lesssim A \|f\|_2$$
\end{proposition}
\begin{proof} This follows from the dual inequality
\begin{equation}\label{dual}
\Big\|\int_1^2 {\mathcal {F}}^{-1}[h\circ(\rho/t)\widehat f_t]\,dt \Big\|_{L^2}
\lesssim A \Big\|\int_1^2 |f_t|dt \Big\|_{L^2(|x|^bdx)}\,.
\end{equation}
Using Plancherel's theorem and generalized polar coordinates
$\xi=\rho\xi'$, $\rho>0$, $\xi'\in \Sigma_\rho$, we can write the
left hand side of \eqref{dual} as
$$\Big(\int \rho^{\frac d\beta-1}\int_{\Sigma_\rho} \Big|\int h(\rho/t)
\widehat f_t(\rho^{1/\beta}\xi')dt \Big| ^2
\frac{d\sigma(\xi')}{\beta |\nabla\rho(\xi')|} d\rho\Big)^{1/2} .
$$
Since $|\nabla \rho|$ is bounded below we can drop this term and use
Lemma \ref{trace} to get
\begin{multline*}
\Big\|\int_1^2 {\mathcal {F}}^{-1}[h\circ(\rho/t)\widehat f_t]dt \Big\|_{L^2}\\
\lesssim
\Big(\int \rho^{\frac d\beta -1}
\int |x|^b \Big|\int h(\rho/t)\rho^{-d/\beta}
f_t(\rho^{-1/\beta}x)dt \Big| ^2 dx\,
d\rho \Big)^{1/2}.
\end{multline*}
We change variables
and the last expression becomes
\begin{align*}
&\Big(\int |x|^b \int \rho^{\frac b\beta-1}
\Big|\int h(\rho/t) f_t(x)dt \Big| ^2 dx\,d\rho\Big)^{1/2}\\
&\le
\Big(\int |x|^b \Big[\int_{1}^2 \Big( \int
| h(\rho/t)|^2 \rho^{\frac b\beta-1} d\rho\Big)^{1/2} | f_t(x)|\,dt
\Big]^2dx\Big)^{1/2}\\
&\le \Big( \int \rho^{\frac b\beta-1}
| h(\rho)|^2d\rho\Big)^{1/2}
\Big(\int
\Big[\int_{1}^2 t^{\frac{b}{\beta}}
| f_t(x)|\,dt \Big]^2|x|^b dx\Big)^{1/2}.
\end{align*}
The assertion follows.
\end{proof}
In the proof of Theorem \ref{sq} we shall use
\begin{proposition} \label{Atauprop}
Let $1<b<d$ and $\rho$ as in Proposition \ref{basicmult}. Let $\eta\in C^\infty$ be supported in $(\frac 18,8)$.
For any real $\tau$ define ${\mathcal {A}}_\tau$ by
\begin{equation}\label{Atau}\widehat {{\mathcal {A}}_\tau f}(\xi) =
\eta(\rho(\xi))
e^{-i\rho(\xi) \tau}
\widehat f(\xi)\,.
\end{equation}
Then
$$\int_{-\infty}^\infty\|{\mathcal {A}}_\tau f\|_{L^2(|x|^{-b})}^2 d\tau \lesssim \|f\|_{2}^2\,.$$
\end{proposition}
\begin{proof}
The inequality is equivalent with
\begin{equation}\label{dualsqf}
\Big\|\int {\mathcal {A}}_\tau [g(\tau, \cdot)] d\tau \Big\|_2 \lesssim
\Big(\int\|g(\tau,\cdot)\|_{L^2(|x|^b)}^2 d\tau\Big)^{1/2}\,.
\end{equation}
Using Plancherel's theorem and generalized polar coordinates we see that
the square of the left hand side is bounded by
\begin{align*}
&\int|\eta(\rho)|^2 \int_{\Sigma_\rho}\Big|
\int e^{i\tau \rho} \widehat g(\tau, \rho^{1/\beta}\xi') d\tau \Big|^2 \frac{d\sigma(\xi')}
{|\nabla\rho(\xi')|} \rho^{\frac d\beta -1}d\rho
\\
&\lesssim \int|\eta(\rho)|^2 \int\Big|
\int e^{i\tau \rho} \rho^{-d/\beta}g(\tau, \rho^{-1/\beta}x)\, d\tau \Big|^2
|x|^{b} dx \, \rho^{\frac d\beta -1} d\rho
\end{align*}
where we have used Lemma \ref{trace}.
By a change of variables the last expression is
\begin{align*}
&\int|\eta(\rho)|^2\rho^{\frac b\beta-1}
\int\Big|
\int e^{i\tau \rho} g(\tau, x) d\tau \Big|^2
|x|^{b} dx \,d\rho
\\
&\lesssim \Big\| \Big(\int \Big|
\int e^{i\tau \rho} g(\tau, \cdot) d\tau \Big|^2 d\rho\Big)^{1/2}
\Big\|_{L^2(|x|^{b})dx}^2
\\
&\lesssim \Big\| \Big(\int |
g(\tau, \cdot)|^2 d\tau \Big)^{1/2}
\Big\|_{L^2(|x|^{b})dx}^2
\end{align*}
where the last inequality holds
by Plancherel's theorem in the first variable.
Now \eqref{dualsqf} follows.
\end{proof}
\subsection*{Littlewood-Paley inequalities}
Let ${\mathfrak {A}}$ be a compact subset of ${\mathbb {R}}^d\setminus \{0\}$ and let $\{\zeta_k\}$ be a bounded family in $C^\infty$, so that all $\zeta_k$ are supported in ${\mathfrak {A}}$.
Define operators $P_k$ by $\widehat {P_k f}(\xi)=\zeta_k(2^{-k} \xi)\widehat f$.
\begin{lemma}\label{LP} Let $-d/2<\gamma<d/2$. Then for all
$f\in\hz{\gamma}{q}$,
$$ \Big\|\Big(\sum_k|P_k f|^2\Big)^{1/2} \Big\|_{\hz{\gamma}{q}} \le C({\mathfrak {A}},\gamma)
\|f\|_{\hz{\gamma}{q}}$$
and
for all $\ell^2$ valued functions
$F=\{f_k\}_{k\in {\mathbb {Z}}} \in \hz{\gamma}{q}(\ell^2)$
$$ \Big\|\sum_kP_k f_k \Big\|_{\hz{\gamma}{q}} \le C({\mathfrak {A}},\gamma)
\Big\|\Big(\sum_k|f_k|^2\Big)^{1/2}\Big\|_{\hz{\gamma}{q}}\,.
$$
\end{lemma}
\begin{proof} Since $|x|^{-2\gamma}$ is an $A_2$ weight for $-d/2<\gamma<d/2$ the Littlewood-Paley inequalities hold for the weighted $L^2(|x|^{2\gamma} dx)$ spaces, i.e. for $\hz{\gamma}{2}$. The case for general $q$ follows by real
interpolation.
\end{proof}
\subsection*{Reduction to the case of homogeneity $\beta=1$}\label{betaeqone}
In the proof of our main theorems we may assume $\beta=1$. This follows
from writing $h(\rho(\xi))= h_\beta(\rho(\xi)^{1/\beta})$ with $h_\beta(s)=h(s^\beta)$ and the following lemma.
\begin{lemma} Let $\alpha>0$, $\beta>0$. Then for all
$h\in B^2_{\alpha, q}({\mathbb {R}})$ supported in a compact
interval $J\subset(0,\infty)$
$$\|h((\cdot)^\beta)\|_{B^2_{\alpha,q}}\le C_{\beta,J }
\|h\|_{B^2_{\alpha,q}}\,.
$$
\end{lemma}
\begin{proof} Let $J=[a,b]$ and $J_\beta=[a^{1/\beta}, b^{1/\beta}]$.
Let $\chi\in C^\infty_0$ be compactly supported in $(0,\infty)$ and equal to $1$ on $J_\beta$. Then we have
$\|\chi(\cdot)h((\cdot)^\beta)\|_{{\mathscr {L}}_k^2}\le C(\chi,\beta)
\|h\|_{B^2_{\alpha,q}}\,
$ for $k=0,1,2,\dots$, by straightforward computation. The corresponding inequality for Besov spaces $B^2_{\alpha,q}$ for all $\alpha>0$
follows by real interpolation.
\end{proof}
\section{Proof of Theorem \ref{herzest}}\label{herzestsect}
In what follows we shall assume $\beta=1$, as we may by the last subsection in \S\ref{prel}.
\subsection*{The basic decompositions}\label{basicdec}
Let $h\in B^2_{\alpha,q'}({\mathbb {R}})$ be supported in $(1/2,2)$ and let $\eta $ be a $C^\infty$ function, supported on $(1/8,8)$ such that
$\eta(s)=1$ on $(1/4,4)$.
Define $L_k$ by
$\widehat {L_k f}(\xi)=\eta(2^{-k}\rho(\xi))\widehat f(\xi)$.
Then
\begin{align}\notag\sup_{s>0} \big|{\mathcal {F}}^{-1}[h(\tfrac \rho s) \widehat f\,]\big|
&=\sup_{k}\sup_{1\le t\le 2} \big|{\mathcal {F}}^{-1}[h(\tfrac{\rho}{2^kt})
\widehat {L_kf}]\big|\,\\
\label{maxsqfct}
&=\Big(\sum_{k}\sup_{1\le t\le 2}
\big|{\mathcal {F}}^{-1}[h(\tfrac{\rho}{2^kt})
\widehat {L_kf}]\big|^2 \Big)^{1/2}\,.
\end{align}
Let $\zeta_0\in C^\infty({\mathbb {R}})$ be supported in $(-1,1)$ such that
$\zeta_0(r)=1$ for $|r|\le 1/2$ and let, for $j\ge 1$,
$\zeta_j(r)=\zeta_0(2^{-j}r)-\zeta_0(2^{-j+1}r)$.
Define
\begin{equation}\label{Vjdef}
V_jh (\rho)=(2\pi)^{-1}\int\zeta_j(r) \widehat h(r) e^{ir\rho}\, dr
\end{equation}
so that $\widehat {V_j h}$ is supported in $I_j=[2^{j-2}, 2^j]\cup [-2^j,-2^{j-2}]$ for $j=1,2,\dots$, and $\sum_{j=0}^\infty V_j h=h$. Define
\begin{equation}\label{Tjtaudef}T^{j,k}_s[h, f]={\mathcal {F}}^{-1}[ \eta(\rho(2^{-k}\cdot))V_jh(2^{-k}
\tfrac{\rho(\cdot)}s)\widehat f\,]\,.\end{equation}
We first note that for large $j$
the Fourier transform of $\eta(\rho(\cdot))V_jh
(\rho(\cdot))$ is concentrated on an annulus of width $\approx
2^j$. By assumption, there is $c_0\in {\mathbb {N}}$ so that \begin{equation}\label{c0def}
2^{-c_0+2}\le |\nabla\rho(\xi)|\le 2^{c_0-2}, \text { for } 1/8\le
\rho(\xi)\le 8. \end{equation}
\begin{lemma} \label{errorlemma}
Let $K_{j,s}$ be defined by $\widehat{K}_{j,s}(\xi)= \eta(\rho(\xi)) V_j h(\rho(\xi)/s)$.
Then for $j=0,1,2,\dots$, and $1/2\le s\le 2$,
\begin{equation}\label{Kjerrors}\begin{aligned}
|K_{j,s}(x)|&\le C_N \|h\|_\infty 2^{-jN}|x|^{-N}, \text{ for }
|x|\ge 2^{j+c_0}\,,
\\
|K_{j,s}(x)|&\le C_N \|h\|_\infty 2^{-jN}, \text{ for }
|x|\le 2^{-j-c_0}\,.
\end{aligned}
\end{equation}
\end{lemma}
\begin{proof}
First observe that $\|\widehat{ V_j h}\|_1\lesssim 2^j\|h\|_\infty$. We write
$$K_{j,s}(x)= \frac{1}{(2\pi)^{d+1}}
\int \widehat {V_jh}(\tau) \int \eta(\rho(\xi))
e^{i(\tau\rho(\xi)+\inn x\xi)}\, d\xi\, d\tau
$$
and observe that, for $\xi \in {\hbox{\roman supp}} (\eta\circ\rho)$,
$$|\nabla_\xi (\tau \rho(\xi)+\inn x\xi)|\ge \max\{
|x|-2^{j+c_0-1}, 2^{j-c_0}-|x|\}$$
Multiple integration by parts in $\xi$ yields the asserted estimate.
\end{proof}
Let $\chi_l$ be the indicator function of the annulus ${\mathfrak {A}}_l$.
The above lemma suggests to
estimate the maximal square-function \eqref{maxsqfct} by
$$\Big(\sum_k {\mathcal {M}}^k_1 [L_k f,h]^2\Big)^{1/2}
+\Big(\sum_k {\mathcal {M}}^k_2 [L_kf,h]^2\Big)^{1/2}
$$
where
\begin{equation} \label{M12def}
\begin{aligned}
{\mathcal {M}}_1^k[ f,h] &= \sup_{1\le t\le 2}\Big|\sum_{j=0}^\infty T^{j,k}_{ t}\big[h,\sum_{n\ge -c_0} \chi\ci{-k+j-n} f\big]\Big|\,,
\\
{\mathcal {M}}_2^k [f,h]&= \sup_{1\le t\le 2}\Big|\sum_{j=0}^\infty T^{j,k}_{ t}\big[h,\sum_{l>j+c_0} \chi\ci{-k+l} f\big]\Big|\,.
\end{aligned}
\end{equation}
For the second term we can reduce to straightforward $L^2$ estimates
for which the weight plays little role.
The first term requires the condition $h\in B^2_{\alpha,s}$ but
we get a bound even for
$\sum_k {\mathcal {M}}_1^k[f_k,h]$ in place of the square-function.
\begin{proposition} \label{M1prop} Let $1/2<\alpha< d/2$, $s=\frac{q}{q-1}$
and let $h\in B^2_{\alpha,s}$ be supported in $(\frac12,2)$.
(i) If $2\le q\le\infty$ then
$$
\Big\|\sum_k {\mathcal {M}}_1^k[f_k,h]
\Big\|_{\hz{-\alpha}{q}} \lesssim \|h\|_{B^2_{\alpha,s}}
\Big\|\Big(\sum_k |f_k|^q\Big)^{1/q}\Big\|_{\hz{-\alpha}{q}}\,.
$$
(ii) If $1\le q\le 2$ then
$$
\Big\|\sum_k {\mathcal {M}}_1^k[f_k,h] \Big\|_{\hz{-\alpha}{2}} \lesssim \|h\|_{B^2_{\alpha,s}}
\Big\|\Big(\sum_k|f_k|^2\Big)^{1/2}\Big\|_{\hz{-\alpha}{q}}\,.
$$
\end{proposition}
\begin{proposition} \label{M2prop} Let $1/2<\alpha< d/2$, $\gamma>1/2$
and let $h\in B^2_{\gamma,1}$ be supported in $(\tfrac 12,2)$. Then
$$
\Big
\|\Big(\sum_k{\mathcal {M}}_2^k[f_k,h]^2\Big)^{1/2}\Big\|_{\hz{-\alpha}{q}}\,
\lesssim \|h\|_{B^2_{\gamma,1}}
\Big\|\Big(\sum_k|f_k|^2\Big)^{1/2}\Big\|_{\hz{-\alpha}{q}}\,.
$$
\end{proposition}
We prove these propositions in the following two subsections.
\subsection*{Estimates for ${\mathcal {M}}_1$} We localize both on the function and the operator sides. The basic estimate is
\begin{lemma} \label{fixed-kmn} Let $1<b<d$ and $1/2<\alpha<d/2$.
Let $V_j$ be as in \eqref{Vjdef}
and
\begin{equation} \label{Lambdajdef}
\Lambda_b^j(h)=
\Big(\int |V_j h(\rho)|^2\rho^{b-1} d\rho\Big)^{1/2}\,.
\end{equation}
For $k\in {\mathbb {Z}}$, $m\in {\mathbb {Z}}$, and $n\ge -c_0$,
\begin{multline*}\Big\|\chi_{m-k} \sup_{1\le t\le 2}\Big|
\sum_{j=0}^\infty T^{j,k}_{t}[h, g\chi\ci{-k+j-n}]\Big| \,\Big\|_{L^2(|x|^{-2\alpha}dx)}
\\
\lesssim 2^{-n\alpha} 2^{m(\frac b2-\alpha)} \sum_{j\ge 0} 2^{j\alpha}\Lambda_b^j(h)
\|g \chi_{j-n-k}\|_{L^2(|x|^{-2\alpha}dx)} \,.
\end{multline*}
\end{lemma}
\begin{proof}
On the support of $\chi_{m-k} $ we have
$|x|^{-2\alpha} \approx 2^{(k-m)(2\alpha-b)}
|x|^{-b}$ and by Minkowski's inequality
\begin{align}\notag
&\Big\|\chi_{m-k} \sup_{1\le t\le 2}\Big|
\sum_{j=0}^\infty T^{j,k}_{ t}[h, g\chi\ci{-k+j-n}]\Big| \,\Big\|_{L^2(|x|^{-2\alpha}dx)}
\\
\label{bpower}
&\lesssim
2^{(k-m)\frac{2\alpha-b}{2}}
\sum_{j=0}^\infty \big\|\sup_{1\le t\le 2}\big| T^{j,k}_{t}[h, g\chi\ci{-k+j-n}]
\big|\,\big\|_{L^2(|x|^{-b}dx)}\,.
\end{align}
Note that
$T^{j,k}_{ t}[h,g](x)= T^{j,0}_{t}[h,g(2^{-k}\cdot)](2^kx)$ and hence
\begin{align*}
&\big\|\sup_{1\le t\le 2}\big| T^{j,k}_{ t}[h, g\chi\ci{-k+j-n}]
\big|\,\big\|_{L^2(|x|^{-b}dx)}\,
\\
&=2^{- k \frac{d-b}2}
\big\|\sup_{1\le t\le 2}\big| T^{j,0}_{t}[h, g(2^{-k\cdot})
\chi\ci{j-n}]
\big|\,\big\|_{L^2(|x|^{-b}dx)}\,
\\&\lesssim 2^{- k \frac{d-b}2} \Lambda_b^j(h) \big\|g(2^{-k}\cdot) \chi_{j-n}\|_2
\end{align*}
where we have applied Proposition \ref{basicmult}. The last displayed quantity is equal to
\begin{align*}
&2^{ kb/2} \Lambda_b^j(h)
\big\|g\chi_{-k+j-n}\|_2
\lesssim
2^{ k\frac{b-2\alpha}2}2^{(j-n)\alpha} \Lambda_b^j(h)
\big\|g\chi_{-k+j-n}\|_{L^2(|y|^{-2\alpha}dy)}
\end{align*}
and the asserted inequality follows if we use this in
\eqref{bpower}.
\end{proof}
\begin{lemma} \label{Lambda-Besov} Let $h\in L^2({\mathbb {R}})$ be supported in $(\tfrac 12, 2)$
and $\Lambda_b^j(h)$ be as in \eqref{Lambdajdef}. Then for $b>0$, $\alpha>0$
$$\Big(\sum_{j=0}^\infty [2^{j\alpha}\Lambda_b^j (h)]^s\Big)^{1/s}\lesssim \|h\|_{B^2_{\alpha,s}}.$$
\end{lemma}
\begin{proof}
Let $$\widetilde \Lambda^j_b(h)\, = \,\int_{1/4}^4|V_j h(\rho)|^2\rho^{b-1}d\rho.$$ Then the corresponding estimate with
$\Lambda^j_b$ replaced by $\widetilde \Lambda^j_b$ is obvious from the definition of Besov spaces. To estimate the corresponding contribution over ${\mathbb {R}}\setminus[1/4,4]$ write
$$ 2\pi V_jh(\rho)= \int h(u)\int\chi_j(r) e^{i (\rho-u)r} dr \,du.$$
By integration by parts, the inner integral is $\le C_N 2^{-jN}|\rho-u|^{-N-1}$
and since
$h$ is supported in $[\tfrac 12,2]$ we see that
$|V_jh(\rho)|\lesssim \|h\|_12^{-jN}$ for $\rho\le 1/4$ and
$|V_jh(\rho)|\lesssim \|h\|_1 2^{-jN}\rho^{-N}$ for $|\rho|\ge 4$.
A straightforward estimate gives the assertion.
\end{proof}
\begin{proof}[Proof of Proposition \ref{M1prop} for $q\ge 2$]
Let
\begin{equation}\label{Mmndef}
{\mathcal {M}}_{m,n}^k f
=\chi_{m-k}
\sup_{1\le t\le 2}\Big|\sum_j T^{j,k}_{t}\big[h,\chi\ci{-k+j-n} f\big]\Big|\,.
\end{equation}
Then, by Minkowski's inequality,
\begin{equation}\label{Mink-mn}
\Big\|\sum_k{\mathcal {M}}_1^k[ f,h]\Big\|_{\hz{-\alpha}{q}}
\le \sum_{m=-\infty}^\infty \sum_{n=-c_0}^\infty
\Big\|\sum_k{\mathcal {M}}_{m,n}^k[ f,h]\Big\|_{\hz{-\alpha}{q}}.
\end{equation}
We use the estimate
\begin{equation}\label{obviousbq}
\Big\|\sum_k \chi_{m-k} g_k\Big\|_{\hz{-\alpha}{q}}
\lesssim \Big(\sum_k \big\|\chi_{m-k}g_k\big\|^q_{L^2(|x|^{-2\alpha})}\Big)^{1/q}
\end{equation}
which holds for all $q\ge 1$ and all $\alpha\in {\mathbb {R}}$ and is
obvious from the definition of the ${\hz{-\alpha}{q}}$ spaces.
We now choose $b_1$, $b_2$ such that $1<b_1<2\alpha$ and $2\alpha<b_2<d$.
From
\eqref {Mink-mn}, \eqref{obviousbq}
and Lemma \ref{fixed-kmn} we get
\begin{multline*}\Big\|\sum_k{\mathcal {M}}_1^k[ f_k,h]\Big\|_{{\hz{-\alpha}{q}}}\,\lesssim\, \sum_{n\ge -c_0}2^{-n\alpha} \,\times \\\Big[
\sum_{m\ge 0} 2^{-m \frac{2\alpha-b_1}2} \Big(\sum_k |{\mathcal {E}}_{b_1,n,k}|^q\Big)^{1/q}
+
\sum_{m<0} 2^{m\frac{b_2-2\alpha}{2}} \Big(\sum_k|{\mathcal {E}}_{b_2,n,k}|^q\Big)^{1/q}\Big]
\end{multline*}
where
\begin{equation}\label{Ebn}{\mathcal {E}}_{b,n,k} \,=\,
\sum_j 2^{j\alpha}\Lambda_b^jh \big\|f_k \chi_{j-n-k}\big\|_{L^2(|y|^{-2\alpha}dy)}
\,.
\end{equation}
Now the proof is concluded by Lemma \ref{Lambda-Besov} and
estimating (with $b=b_1$ or $b=b_2$)
\begin{align*}
&\Big(\sum_k|{\mathcal {E}}_{b,n,k}|^q\Big)^{1/q}\\&\lesssim
\Big(\sum_k
\Big(\sum_j [2^{j\alpha}\Lambda_b^j(h)]^s\Big)^{q/s}
\sum_j
\big\|f_k \chi_{j-n-k}\big\|_{L^2(|y|^{-2\alpha}dy)}^q\Big)^{1/q}
\\&=\Big(\sum_j [2^{j\alpha}\Lambda_b^j(h)]^s\Big)^{1/s}
\Big(\sum_l\sum_k
\big\|f_k \chi_{l}\big\|_{L^2(|y|^{-2\alpha})}^q\Big)^{1/q}
\\&\lesssim \|h\|_{B^2_{\alpha,s}}
\Big\|\Big(\sum_k |f_k|^q\Big)^{1/q}\Big\|_{\hz{-\alpha}{q}}\,. \qedhere
\end{align*}
\end{proof}
\begin{proof}[Proof of Proposition \ref{M1prop} for $q\le 2$]
Again we work with \eqref{Mmndef}, use \eqref{Mink-mn}, \eqref{obviousbq}
for $q=2$ and Lemma \ref{fixed-kmn} to get
\begin{multline*}\Big\|\sum_k{\mathcal {M}}_1^k[ f_k,h]\Big\|_{{\hz{-\alpha}{2}}}\,\lesssim\, \sum_{n\ge -c_0}2^{-n\alpha} \,\times \\\Big[
\sum_{m\ge 0} 2^{-m \frac{2\alpha-b_1}2} \Big(\sum_k |{\mathcal {E}}_{b_1,n,k}|^2\Big)^{1/2}
+
\sum_{m<0} 2^{m\frac{b_2-2\alpha}{2}} \Big(\sum_k|{\mathcal {E}}_{b_2,n,k}|^2\Big)^{1/2}\Big],
\end{multline*}
with ${\mathcal {E}}_{b,n,k}$ as in \eqref{Ebn}, and
$b_1\in (1,2\alpha)$ and $b_2\in(2\alpha,d)$.
Again we apply H\"older's inequality and get, for fixed $m, n$,
\begin{align*}
&\Big(\sum_k|{\mathcal {E}}_{b,n,k}|^2\Big)^{1/2}
\\
&\le\Big(\sum_k
\Big(\sum_j [2^{j\alpha}\Lambda_b^j(h)]^s\Big)^{2/s}\Big(\sum_j \big\|f_k \chi_{j-n-k}\big\|_{L^2(|y|^{-2\alpha}dy)}^q\Big)^{2/q}\Big)^{1/2}
\\
&\le \Big(\sum_j [2^{j\alpha}\Lambda_b^j(h)]^s\Big)^{1/s}
\Big(\sum_k\Big[\sum_l \big\|f_k \chi_{l}\big\|_{L^2(|y|^{-2\alpha}dy)}^q\Big]^{2/q}
\Big)^{1/2}
\end{align*} and since now $q\le 2$ we may use Minkowski's inequality
and estimate
this by
\begin{align*}
&\|h\|_{B^2_{\alpha,s}} \Big(\sum_l \Big\|\chi_l\Big(\sum_k|f_k|^2\Big)^{1/2}
\Big\|_{L^2(|y|^{-2\alpha}dy)}^q\Big)^{1/q}
\\
&\lesssim
\|h\|_{B^2_{\alpha,s}} \Big\|\Big(\sum_k|f_k|^2\Big)^{1/2}
\Big\|_{\hz{-\alpha}{q}}\,. \qedhere
\end{align*}
\end{proof}
\subsection*{Estimates for ${\mathcal {M}}_2$}\label{m2sect}
For every $l$ let
$\widetilde\chi\ci{l}= \sum_{i=-2-c_0}^{2+c_0} \chi_l$.
We estimate
$${\mathcal {M}}_2^k[f,h] \le \sum_{j\ge 1} \Big(\widetilde {\mathcal {M}}_{2,j}^k[f,h]
+{\mathcal {N}}_j^k[f,h]\Big)\,,$$ where \begin{equation} \label{M2defs}
\begin{aligned}
\widetilde {\mathcal {M}}_{2,j}^k[ f,h] &= \sum_{l>j+c_0} \sup_{1\le t\le 2}\Big| \widetilde \chi\ci{-k+l}
T^j_{2^k t}\big[h, \chi\ci{-k+l} f\big]\Big|\,,
\\
{\mathcal {N}}_j^k [f,h]&= \sum_{l>j+c_0}\sup_{1\le t\le 2}\Big|(1-\widetilde \chi\ci{-k+l})\sum_j T^j_{2^k t}\big[h,
\chi\ci{-k+l} f\big]\Big|\,.
\end{aligned}
\end{equation}
For the estimation of $(\sum_k\widetilde {\mathcal {M}}_{2,j}^k[f,h] ^2)^{1/2}$
we use a standard imbedding estimate.
\begin{lemma}\label{Sobemb}
For $\gamma>1/2$
$$\big\|\sup_{1\le t\le 2}\big|T_{2^kt}^j [h,f]\big|\big\|_2
\le 2^{j(\frac 12 -\gamma)} \| h\|_{B^2_{\gamma,1} } \|f\|_{L^2({\mathbb {R}}^d)}.$$
\end{lemma}
\begin{proof}
By scaling it suffices to prove this for $k=0$.
For $g$ supported in $(1/2,2)$ and $I=[1/2,2]$,
$$\|\sup_{t\in I}|{\mathcal {F}}^{-1}[g(\tfrac{\rho(\cdot)}{t})\widehat f\,]|\big\|_2 \le
\big[ \|g\|_2 + \|g\|_2^{1/2}\|g'\|_2^{1/2}\big] \|f\|_2\,,
$$
as one can see by applying the fundamental theorem of calculus to $g^2$. Since
$\|V_j h\|_2 + 2^{-j}\|(V_j h)'\|_2 \le
2^{-j\gamma} \|h\|_{B^2_{\gamma,1}}$,
the assertion follows.
\end{proof}
\begin{proof}[Proof of Proposition \ref{M2prop}]
From Lemma \ref{Sobemb} we get, considering the supports of the
functions $\chi\ci{-k+l}$ and $\widetilde \chi\ci{-k+l}$,
\begin{align*}
\\ &\Big\| \Big(\sum_k \widetilde {\mathcal {M}}_{2,j}^k[ f_k,h]^2\Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha} dx)}
\\
&\lesssim \Big(\sum_{l> c_0+j}\sum_k2^{2\alpha(k-l)}
\big\|\sup_{1\le t\le 2}\big|
T^{j,k}_{t}\big[h, \chi\ci{-k+l} f_k ]\big|\big\|_2^2 \Big)^{1/2}
\\
&\lesssim
2^{j(\frac 12-\gamma)}\|h\|_{B^2_{\gamma,1}}
\Big(\sum_{l}\sum_k2^{2\alpha(k-l)}
\big\|f_k\chi\ci{-k+l}\big\|_2^2\Big)^{1/2}
\\&\lesssim 2^{j(\frac 12-\gamma)}\|h\|_{B^2_{\gamma,1}}
\Big\|\Big(\sum_k|f_k|^2\Big)^{1/2}\Big\|_{L^2(|x|^{-2\alpha}dx)}\,.
\end{align*}
From Lemma \ref{errorlemma} we get
$${\mathcal {N}}_{j}^k[ f,h](x) \lesssim 2^{-jN} \|h\|_\infty \int
\frac{2^{kd}}{(1+2^k|x-y|)^N}
|f(y)|\, dy\,,
$$
and since $|x|^{-2\alpha}$ is an $A_2$ weight for $|\alpha|<d/2$ the estimate
$$\Big\| \Big(\sum_k {\mathcal {N}}_{j}^k[ f_k,h]^2\Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha}dx)}
\lesssim 2^{-jN}\|h\|_\infty
\Big\|\Big(\sum_k|f_k|^2\Big)^{1/2}\Big\|_{L^2(|x|^{-2\alpha}dx)}
$$
follows from standard results (see ch. IV in \cite{g-r}).
Proposition
\ref{M2prop} follows by combining the estimates for $\{\widetilde{\mathcal {M}}^k_{2,j}\}$ and $\{{\mathcal {N}}^k_j\}$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{herzest}] By \eqref{maxsqfct}, the decomposition described in \S\ref{basicdec} and
Propositions \ref{M1prop} and \ref{M2prop} we have the estimates
$$\Big\|\sup_{t>0}|{\mathcal {F}}^{-1}[h(\tfrac{\rho(\cdot)}{t})\widehat f\,]|
\Big\|_{\hz{-\alpha}{\sigma}} \lesssim
\Big\|\Big(\sum_k|P_k f|^2\Big)^{1/2}\Big\|_{\hz{-\alpha}{q}}, \text{ $\sigma=\max\{q,2\}$,}
$$ where
the $P_k$ are suitable Littlewood-Paley operators as used in Lemma \ref{LP}, satisfying $L_kP_k=L_k$. We conclude by an application of
the first inequality in Lemma \ref{LP}.
\end{proof}
\section{Conclusion of other proofs}\label{concl}
\begin{proof}[Proof of Theorem \ref{multch}] $(i)$ and $(ii)$ are equivalent by duality.
Assume $(ii)$ holds. Define $T_t $ by $\widehat{ T_tf}(\xi)=
m(t\rho(\xi))
\widehat f(\xi)$.
Then $T_t$ is bounded on $L^2(|x|^{2\alpha})=\hz{\alpha}{2}$, with operator norm independent of $t$,
by the homogeneity of $\rho$ and scale invariance. Test $T_t$ on Schwartz functions of the form $\varphi\circ \rho$
to derive the condition $(iii)$.
Finally if $(iii)$ holds, let $\chi\in C^\infty({\mathbb {R}})$ be supported
in $(1/2,2)$ so that $\sum_{k\in{\mathbb {Z}}}[\chi(2^{-k}s)]^3=1$. Define
$P_k$ and $T_k$ by $\widehat {P_k
f}(\xi)=\chi(2^{-k}\rho(\xi))\widehat f(\xi)$ and $\widehat {T_k
f}(\xi)=\chi(2^{-k}\rho(\xi))m(\rho(\xi))\widehat f(\xi)$.
It follows from Propositions
\ref{M1prop} and \ref{M2prop} that
$P_kT_kP_k$ is bounded on $\hz{-\alpha}{2}$ with operator norm $\lesssim
\|\chi m(2^k\cdot)\|_{{\mathscr {L}}^2_\alpha}$. Hence by interchanging sums and integrals
$$\Big\|\Big(\sum_k|P_kT_kP_k f|^2\Big)^{1/2}\Big\|_{\hz{-\alpha}{2}}
\lesssim \sup_k \|\chi m(2^k\cdot)\|_{{\mathscr {L}}^2_\alpha}
\Big\|\Big(\sum_k|P_kf|^2\Big)^{1/2}\Big\|_{\hz{-\alpha}{2}}. $$
A standard estimation gives
$\sup_k \|\chi m(2^k\cdot)\|_{{\mathscr {L}}^2_\alpha}
\le \sup_t \|\varphi m(t\cdot)\|_{{\mathscr {L}}^2_\alpha} $.
We finish by applying both Littlewood-Paley inequalities in Lemma \ref{LP}.
\end{proof}
\begin{proof}[Proof of Corollary \ref{herzcor}]
Given Theorem \ref{herzest} this is a standard argument.
Define $S_t$ by
$\widehat {S_tf}(\xi)=\chi(\rho(\xi)/t)h(\rho(\xi)/t)$ with
$h\in B^2_{\alpha, q'}$.
Let $f\in L^{p,q}$. We need to show
$\lim_{t\to\infty}S_t f(x)=0$ a.e. on $U_R=\{x:R^{-1}\le |x|\le R\}$, for every $R>2$. For this it suffices to show that for every $r>0$, $\varepsilon>0$ we have
\begin{equation} {\text{\rm meas}}\{x\in U_R: \limsup_{t\to \infty}|S_t f(x)|>r\}<\varepsilon\,.
\label{limsup}
\end{equation}
It is easy to see that for any Schwartz function $g$
we have
$\lim_{t\to\infty}S_t g(x)=0$ uniformly on compact sets.
Let $\sigma=\max\{q,2\}$.
Given any Schwartz function $g$ the left hand side of \eqref{limsup} can now be estimated by
\begin{align*}
&{\text{\rm meas}}\{x\in U_R: \limsup_{t\to \infty}|S_t [f-g](x)|>r\}
\\ &\lesssim r^{-2} \| \sup_{t>0} |S_t [f-g]|\|_{L^2(U_R)}
\\ &\lesssim r^{-2} R^{d(1/2-1/p)}(\log R)^{1/2-1/\sigma}\| \sup_{t>0}
|S_t [f-g]|\|_{\hz{d(\frac12-\frac 1p)} {q}}\,.
\end{align*}
If we combine Theorem \ref{herzest} with Lemma \ref{Loremb} we get
$$\| \sup_{t>0}
|S_t [f-g]|\|_{\hz{d(\frac12-\frac 1p)} {q}}\,\lesssim \|h_{\lambda} \def\La{\Lambda,\gamma}\|_{B^2_{d(1/2-1/q),q'}}
\|f-g\|_{L^{p,q}}\,.
$$
The implicit constants in these estimates are independent of $g$.
Since ${\mathcal {S}}$ is dense in $L^{p,q}$ we may find $g\in {\mathcal {S}}$ to make the last
expression as small as we wish, and \eqref{limsup} follows. \end{proof}
\begin{proof}[Proof of Theorem \ref{laga}]\
Let $\chi\in C^\infty_0$ supported in $(1/2,2)$ and equal to $1$ in a neighborhood of $1$. We have $h_{\lambda} \def\La{\Lambda,\gamma}(\rho/t)\widehat f(\xi)=
\widehat T_t f(\xi)+ \widehat S_t f(\xi)$ where
$T_t$ is defined by by
$\widehat {T_tf}(\xi)=(1-\chi(\rho(\xi)/t)) h_{\lambda} \def\La{\Lambda,\gamma}(\rho(\xi)/t)
\widehat f(\xi)$. The maximal function $\sup_t|T_tf|$ is controlled by the Hardy-Littlewood maximal function and it is standard that $\lim T_t f(x)=f(x)$ almost everywhere for every $f\in L^r$, $r\ge 1$.
Next
$h_{\lambda} \def\La{\Lambda,\gamma} \in B^2_{\alpha, q'}$
for
$\alpha=d(1/2-1/p)$ and $\gamma>1-1/q$, moreover
$h_{\lambda} \def\La{\Lambda,0} \in B^2_{\alpha, \infty}$.
By Corollary \ref{herzcor}, $S_t f(x)\to 0$ a.e. for $f\in
L^{p,q}$. This proves the theorem.
\end{proof}
\section{The square-function estimate}\label{sqsect}
We now give a proof of Theorem \ref{sq}.
Let $\rho \in C^\infty_0({\mathbb {R}}^d)$ be homogeneous of degree $1$, with $\rho(\xi)>0$ for $\xi\neq 0$.
We aim to prove, for $1/2<\alpha<d/2$,
\begin{equation}\label{Galphabeta}
\Big\|\Big(\int_0^\infty \big|{\mathcal {F}}^{-1}
\big[\frac{\rho^\beta}{t^\beta}
(1-\frac{\rho^\beta}{t^\beta})_+^{\alpha-1} \widehat f
\,\big]\big|^2 \frac{dt}{t}\Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha}dx)} \lesssim \|f\|_{L^2(|x|^{-2\alpha}dx)}\,.
\end{equation}
A change of variables shows that \eqref{Galphabeta} implies
Theorem \ref{sq}. Let $\eta_0\in C^\infty_0$ be supported on $(-5/6,5/6)$ and
equal to $1$ on $(-3/4,3/4)$. By standard weighted norm estimates for
vector-valued singular integrals we have
$$
\Big\|\Big(\int_0^\infty \big|{\mathcal {F}}^{-1} \big[\eta_0(\frac\rho t)
\frac{\rho^\beta}{t^\beta}
(1-\frac{\rho^\beta}{t^\beta})_+^{\alpha-1} \widehat f
\,\big]\big|^2 \frac{dt}{t}\Big)^{1/2} \Big\|_{L^2(|x|^{b}dx)} \lesssim
\|f\|_{L^2(|x|^{b}dx)}
$$
for all $\alpha>0$, $\beta>0$ and $-d<b<d$. Thus it suffices to prove
$$
\Big\|\Big(\int_0^\infty \big|{\mathcal {F}}^{-1} \big[\eta_1(\frac \rho t)
(1-\frac{\rho^\beta}{t^\beta})_+^{\alpha-1}
\widehat f \,\big]\big|^2 \frac{dt}{t}\Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha}dx)}
\lesssim \|f\|_{L^2(|x|^{-2\alpha}dx)}
$$
for $\eta_1\in C^\infty$ supported in $ (3/4,5/4)$.
By Littlewood-Paley theory for the spaces $L^2(|x|^{-2\alpha}dx)$ and scaling it suffices to prove the analogous inequality
in the $t$-localized case for which in the last display
the integral over ${\mathbb {R}}^+$ is replaced by the integral over $[1,2]$.
We observe also that, with $\eta_2 \in C^\infty_c(0,\infty)$,
$$
\frac { \eta_2(\tfrac {\rho(\xi)}{t})(1-\tfrac{\rho(\xi)^\beta}{t^\beta})_+^{\alpha-1}}
{ (1-\tfrac{\rho(\xi)}{t})_+^{\alpha-1}} = g_\alpha(\xi/t)
$$
where $g_\alpha\in {\mathscr {L}}^2_\beta$ for all $\alpha>0$, $\beta>0$. Thus,
by an elementary inequality for vector valued operators
it suffices to prove that for $1/2<\alpha<d/2$
$$
\Big\|\Big(\int \big|{\mathcal {F}}^{-1} \big[\eta(\rho)
\varsigma(t)
(t-\rho)_+^{\alpha-1}
\widehat f \,\big]\big|^2 dt\Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha}dx)}
\lesssim \|f\|_{L^2(|x|^{-2\alpha}dx)}\,;
$$
here $\varsigma$ is a compactly supported $C^\infty$ function.
By applying Plancherel's theorem in $t$ the preceding inequality
follows from
\begin{equation}\label{GalphalocF}
\Big\|\Big(\int_{-\infty}^\infty (1+|\tau|^2)^{-\alpha}|{\mathcal {A}}_\tau f|^2 d\tau \Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha}dx)}
\lesssim \|f\|_{L^2(|x|^{-2\alpha}dx)}
\end{equation}
with ${\mathcal {A}}_\tau $ as in \eqref{Atau}
As before let $\chi_l$ be the characteristic function of the annulus ${\mathfrak {A}}_l$.
Let $I_0=[-1,1]$ and, for $j\ge 1$, $I_j=[-2^{j}, -2^{j-1}]\cup[2^{j-1}, 2^j]$.
In order to estimate
\eqref{GalphalocF}
we make a decomposition which is analogous to the one in \eqref{M12def}, and prove
\begin{multline}
\label{hardytype}
\Big\|\Big(\sum_{j=0}^\infty 2^{-2j\alpha}
\int_{I_j}\big|{\mathcal {A}}_\tau (\sum_{n\ge -c_0} f\chi_{j-n})\big|^2 d\tau \Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha}dx)}\\ \lesssim
\|f\|_{L^2(|x|^{-2\alpha}dx)}\,,
\end{multline} and \begin{multline}
\label{L2type}
\Big\|\Big(\sum_{j=0}^\infty 2^{-2j\alpha}
\int_{I_j}\big|{\mathcal {A}}_\tau (\sum_{l\ge j+c_0} f\chi_{l})\big|^2 d\tau \Big)^{1/2}
\Big\|_{L^2(|x|^{-2\alpha}dx)} \\ \lesssim
\|f\|_{L^2(|x|^{-2\alpha}dx)}\,.
\end{multline}
In order to show \eqref{hardytype} we prove for $m\in {\mathbb {Z}}$, $n\ge -c_0$,
$\varepsilon <\min\{\alpha-\tfrac 12, \tfrac d2-\alpha\}$,
\begin{equation}\label{Gmn} \Big(\sum_{j=1}^\infty2^{-2j\alpha}\int_{I_j}
\big\|\chi_m {\mathcal {A}}_\tau(f\chi_{j-n})
\big\|_{L^2(|x|^{-2\alpha}dx)}^2d\tau\Big)^{1/2} \lesssim 2^{-|m|\varepsilon}
2^{-n\alpha} \|f\|_2\,. \end{equation}
For the proof of \eqref{Gmn}
let $1<b<d$. The left hand side of \eqref{Gmn} can be estimated by
\begin{align*}& 2^{-m\frac{2\alpha -b}{2}} \Big(\sum_j 2^{-2j\alpha}
\int
\big\|{\mathcal {A}}_\tau (f\chi\ci{j-n})\big\|_{L^2(|x|^{-b}dx)}^2 d\tau\Big)^{1/2}
\\
&\lesssim 2^{-m\frac{2\alpha -b}{2}} \Big(\sum_j 2^{-2j\alpha}
\big\|f\chi\ci{j-n}\big\|_{2}^2 \Big)^{1/2}\,,
\end{align*} by Proposition \ref{Atauprop}. The last display is estimated by
\begin{align*} & 2^{-m\frac{2\alpha -b}{2}} 2^{-n\alpha} \Big(\sum_j
\big\|f\chi\ci{j-n}\big\|_{L^2(|x|^{-2\alpha}dx)}^2 \Big)^{1/2}
\\
&\lesssim 2^{-m\frac{2\alpha -b}{2}} 2^{-n\alpha} \|f\|_{L^2(|x|^{-2\alpha}dx)}\,.
\end{align*}
We may choose $b$ such that
$1<b<2(\alpha-\varepsilon)$ if $m>0$ and $2(\alpha+\varepsilon)<b<d$ if $m\le 0$ and
then \eqref{Gmn} follows.
We sketch the proof of the more straightforward inequality
\eqref{L2type} and rely on the argument used for ${\mathcal {M}}^2_k$ before.
Let $\widetilde \chi_l$ be the characteristic function of
$\cup_{-5\le i\le 5} {\mathfrak {A}}_{l+i}$. Then for $l>j+c_0$ and $\tau\in
I_j$
$$(1-\widetilde \chi_{l}){\mathcal {A}}_\tau(f\chi_l)
(x)\le C_N 2^{-lN}
\int\frac{|f(y)\chi_l(y)|}{
(1+2^l |x-y|)^{N} }dy
\,.$$ Since
$|x|^{-2\alpha}$ is an $A_2$ weight, it is immediate that
$$\Big(\sum_{j=0}^\infty 2^{-2j\alpha}\int_{I_j}
\Big\|\sum_{l>j} (1-\widetilde \chi_{l})|{\mathcal {A}}_\tau(f\chi_l)|
\Big\|_{L^2(|x|^{-2\alpha}dx)}^2 d\tau \Big)^{1/2}
\lesssim
\|f\|_{L^2(|x|^{-2\alpha}dx)}.
$$
For the main term we use $\alpha>1/2$ and estimate
\begin{align*}
&\Big(\sum_{j=0}^\infty 2^{-2j\alpha}\int_{I_j}
\Big\|\sum_{l>j} \widetilde \chi_{l}|{\mathcal {A}}_\tau(f\chi_l)|
\Big\|_{L^2(|x|^{-2\alpha}dx)}^2d\tau\Big)^{1/2}
\\ &\lesssim
\Big(\sum_{j=0}^\infty 2^{-2j\alpha}\int_{I_j}\sum_{l>j}2^{-2l\alpha}
\|{\mathcal {A}}_\tau(f\chi_l)\|_2^2\,d\tau
\Big)^{1/2}
\\ &\lesssim
\Big(\sum_{l>0}\sum_{j=0}^{l-1} 2^{-2j(\alpha-1)}2^{-2l\alpha}\|f\chi_l\|_2^2
\Big)^{1/2} \lesssim
\|f\|_{L^2(|x|^{-2\alpha}dx)}.
\end{align*}
Thus \eqref{L2type} is proved.\qed
|
\section{Introduction}
\label{sec:intro}
The production of quarks and gluons via strong interactions is the dominant high-momentum-transfer
process
at the LHC and is a significant background to most new-physics searches.
These partons are measured as jets, which are collimated streams of charged and neutral particles, clustered using dedicated algorithms.
Corrections to measured quantities are necessary to relate the jets to their
parent partons. Many gluons are generated in most common Standard Model processes, such as the inclusive production of
jets~\cite{InclJets,Aad:2011tqa}. On the other hand,
some processes arising from new-physics models, for example supersymmetry, generate many light quarks~\cite{ATLASSUSY1,ATLASSUSY2}.
The power to discriminate between jets initiated by light quarks and those initiated
by gluons would therefore provide a powerful tool in searches for new physics.
In case of a discovery of a new particle, such a discriminant could provide
valuable information about its nature.
Also, some Standard Model measurements rely on the correct identification of the origin of jets,
as in the cases of reconstructing a hadronic $\Wboson$ decay when measuring the top quark mass, or in the reconstruction
of a hadronic $\ensuremath{Z}$ decay when measuring the Higgs boson mass via $h\rightarrow ZZ\rightarrow\ell\ell \antibar{q}$.
These analyses would benefit from such a discriminant. These applications motivate the analysis of the partonic
origin of jets that is the focus of this paper.
In perturbative quantum chromodynamics (QCD), the concept of a parton initiating a
jet is a fixed-order notion. In the matrix-element
calculation of a high-momentum-transfer-process, the outgoing partons appear na\"ively much like outgoing
particles in the final state. However, only colourless states with two or more partons can
form an observable jet. Moreover, in a parton shower, the leading parton is only
well defined for a fixed number of splittings. The next step in the shower may change
the energy, direction, or flavour of the leading parton. Thus, labelling jets with a specific flavour
and interpreting results after such labelling requires a clearly defined procedure~\cite{schwartz1}.
Certain parton branchings can yield an ambiguous jet identity.
The labelling of a jet may also depend on the physics goal of the analysis.
For example, a jet from the $\antibar{q}'$ decay of a high-momentum $\Wboson$ boson produced in
a top quark decay can be considered either as a part of a top-quark jet or as a boosted
$\Wboson$-boson jet.
Nonetheless, many event topologies lend themselves to the identification of a jet as having
originated from a specific type of parton in the matrix-element calculation. Such an approach
can lead to an unambiguous and meaningful parton labelling for a large majority of jets.
This approach of linking jet-by-jet labelling to the results of the underlying leading-order (LO)
calculation is also used in this paper to define the flavour of a jet.
Discrimination between jets of different partonic origin has been attempted previously at several
experiments~\cite{TevatronShapes1,QGNN,Pumplin,QGopal,QGsub,QGlep,QGcleo,DelphiQG,DelphiQG2,AlephQG,L3QG}.
Most work has relied on jet properties that result from the difference in colour charge between
the partons. The colour factors in quantum chromodynamics differ for quarks ($C_F=4/3$) and gluons
($C_A=3$), and therefore, for example, one expects approximately $C_A/C_F=9/4$ times more particles
in a gluon-initiated jet than in a jet initiated by a light ($u$, $d$ or $s$) quark.
The measured difference in particle multiplicity at OPAL was, in fact, not far from this expectation~\cite{QGopal}.
Because of the showering that produces these additional particles, gluon jets are also expected
to be wider and have a softer particle spectrum.
The most successful studies of discrimination between light-quark-initiated and gluon-initiated jets
(henceforth, quark-jets and gluon-jets) have taken place at electron-positron colliders~\cite{QGinee1,QGinee2}.
The selection and identification of ``pure'' samples of quark- and
gluon-jets is considerably more difficult at hadron colliders because of the complication added
by beam remnants, initial-state radiation, and multi-parton interactions. The presence
of multiple soft $pp$ collisions overlaying the hard-scatter interaction of interest at the LHC further
complicates this task.
Recently some effort has been devoted to developing kinematic selections that significantly
enhance the fraction of quark-jets or gluon-jets in a set of events~\cite{schwartz1}.
In addition, discriminants based on jet structure have shown some promise for
distinguishing between classes of jets at the LHC~\cite{schwartz2}.
Jets that include, or are initiated by, heavy quarks (bottom and charm) also exhibit properties
different from those of quark-jets~\cite{TevatronHQ,ATLASbjetshapes}. Generally, these jets are wider than
quark-jets. They are often identified by long-lived or leptonically decaying hadrons.
However, no special discriminant for them is developed here.
This paper is organised as follows. The ATLAS detector is briefly described in Sec.~\ref{sec:atlas}.
Section~\ref{sec:data-selection} describes details of the data and Monte Carlo (MC) samples
used, as well as the object reconstruction and event selection. Section~\ref{sec:jetdef} introduces
the definition of gluon-jets and quark-jets that are used in the remainder of the paper.
The jet properties used to build a discriminant from samples with different purities,
and the validation of the extraction method
using MC event samples, are described in Sec.~\ref{sec:mcTemp}.
Section~\ref{sec:disc} describes the selection of samples based on kinematic variables
to enhance quark-jet or gluon-jet fractions and the validation of the extracted properties using
those samples.
The likelihood-based discriminant is described in Sec.~\ref{sec:perfDetails}, where its performance
in MC simulation and in data is discussed.
Finally, the conclusions are presented in Sec.~\ref{sec:conclusions}.
\section{ATLAS detector}
\label{sec:atlas}
The ATLAS detector~\cite{techpaper} comprises an inner tracking detector, a calorimeter system, and
a muon spectrometer.
The inner detector (ID) includes a silicon pixel detector, a silicon microstrip detector and a transition radiation tracker.
It is immersed in a 2~T axial magnetic field provided by a solenoid and precisely measures the trajectories of
charged particles with $|\eta|<2.5$~\footnote{ATLAS uses a right-handed coordinate system with its
origin at the nominal interaction point (IP) in the centre of the detector and the $z$-axis along the
beam pipe.
The $x$-axis points from the IP to the centre of the LHC ring, and the $y$-axis points upward.
Cylindrical coordinates $(r,\phi)$ are used in the transverse plane, $\phi$ being the azimuthal angle
around the beam pipe.
The pseudorapidity is defined in terms of the polar angle $\theta$ as $\eta=-\ln\tan(\theta/2)$, and
the rapidity is defined as $\frac{1}{2}\ln{\left(\frac{E+p_Z}{E-p_Z}\right)}$, where $E$ is the object's energy and $p_z$ is
its momentum along the $z$-axis. The values of $\eta$, $\phi$, and $y$ are determined at the interaction vertex.
The variable $\ensuremath{\Delta R}\xspace=\sqrt{(\Delta\eta)^2+(\Delta\phi)^2}$ is used to characterise the
angular difference between two objects using their $\eta$ and $\phi$ directions.}.
The calorimeter system covers the region $|\eta|<4.9$ and is divided into electromagnetic and hadronic
compartments.
Electromagnetic calorimetry in the region $|\eta|<3.2$ is provided by liquid-argon
sampling calorimeters with lead absorbers.
In the barrel region ($|\eta|<1.7$), the hadronic calorimeter comprises scintillator tiles with steel
absorbers, and the endcap region ($1.4<|\eta|<3.2$) is covered by a liquid-argon and copper
sampling hadronic calorimeter.
The calorimetry in the forward region ($3.2<|\eta|<4.9$) is provided by a liquid-argon and copper sampling
electromagnetic calorimeter and a liquid-argon and tungsten sampling hadronic calorimeter.
The muon spectrometer (MS) covers $|\eta|<2.7$ and uses a system of air-core toroidal magnets.
ATLAS has a three-level trigger system to select events.
The first-level trigger uses custom-built hardware components and
identifies jet, electron and photon candidates using coarse calorimeter information,
and muon candidates using coarse tracking information from the muon spectrometer.
At the highest level, full event reconstruction, similar to that used in the offline software, is performed
to accurately identify and measure objects that determine whether the event is recorded.
\section{Data sample and event selection}
\label{sec:data-selection}
Several samples are used in the construction and validation of the variables entering the
quark/gluon discriminant: dijet events, trijet events, $\gamma$+jet events,
$\gamma$+2-jet events, $t\bar{t}$ events and $W$+jet events.
After basic data quality requirements are imposed to remove known detector errors and readout problems,
the selected dataset corresponds to a total integrated luminosity of 4.67$\pm$0.08~fb$^{-1}$~\cite{NewLumiPaper}.
The data were collected from March to October 2011
at a centre-of-mass energy $\sqrt{s}=7\TeV$.
The average number of additional $pp$ collisions per bunch crossing, called
``pile-up'', rose during the data-taking period from a few to 15.
\subsection{Monte Carlo simulation}\label{sec:mc}
Simulated event samples are generated for comparison with data and
for the determination of the systematic uncertainties based on variations in
the MC generator settings. For the MC samples, several different generators are used.
{\sc MadGraph}~\cite{MadGraph} is run as a $2\rightarrow N$ generator with MLM matching~\cite{MLM}, uses
the CTEQ 6L1 parton distribution function (PDF) set, and is interfaced to {\sc Pythia} 6 with a version
of the ATLAS {\sc MC11} Underlying Event Tune 2B ({\sc AUET2B})~\cite{ATL-PHYS-PUB-2011-009}
constructed for this PDF set. {\sc Herwig}++~\cite{Herwigpp} is run standalone
as a $2\rightarrow 2$ generator and uses the MRST LO** PDF set with the {\sc LHC-UE7-2}
tune~\cite{Gieseke:2011xy}. This tune of {\sc Herwig}++ has an improved description
of colour reconnection in multiple parton interactions and has been shown to have fair agreement
with ATLAS data in minimum-bias observables~\cite{Gieseke:2011xy}.
{\sc Pythia} 6 is also run standalone as a $2\rightarrow 2$ generator with the MRST LO**
PDF set and the {\sc AUET2B} tune.
The {\sc AUET2B} tune incorporates ATLAS~\cite{ATLASshape} and CDF~\cite{CDFshape} jet-shape measurements as
well as ATLAS fragmentation function measurements at $\sqrt{s}=7$~\TeV~\cite{ATLASfrag} and is thus expected
to describe inclusive-jet properties well.
Additional pile-up events, which are superimposed on the hard-scattering event, are generated with either
{\sc Pythia} 6~\cite{pythia} with the {\sc AUET2B} tune using the MRST LO** PDF~\cite{Martin:2009iq}
set, or {\sc Pythia} 8~\cite{Sjostrand:2007gs} with the {\sc 4C} tune~\cite{4CTune} using the CTEQ 6L1 PDF set~\cite{Pumplin:2002vw}.
Choosing between these two pile-up simulations has negligible impact on the analysis. The
number of pile-up events in the MC simulation is reweighted to match the conditions
found in the data for each trigger selection. The events are passed through the ATLAS detector
simulation~\cite{simulation}, based on {\sc GEANT}4\cite{geant}, and are reconstructed using the same
software as for the data.
\subsection{Jet reconstruction, selection and calibration}
Jets are constructed from topological clusters of calorimeter cells~\cite{Lampl:1099735} and
calibrated using the EM+JES scheme~\cite{jesPaper,InclJets}.
This scheme is designed to adjust the energy measured in the calorimeter to that of the
true particle jets on average.
Calorimeter jets are reconstructed using
the anti-\ensuremath{k_{t}}\xspace jet algorithm~\cite{Cacciari:2008gp,FastJet} with a four-momentum recombination scheme
and studied if calibrated transverse momentum $\pt>20$~\GeV\ and $|\eta|<4.5$.
Jet-finding radius parameters of both $R=0.4$ and $R=0.6$ are studied. Only jets with
$|\eta|<2.1$ are used for building the quark-jet tagger, to guarantee that the jet is well within
the tracking acceptance.
In the MC simulation, particle jets are reconstructed using the same anti-\ensuremath{k_{t}}\xspace algorithm with stable, interacting particles~\footnote{
A particle is considered stable and interacting if its lifetime is longer than 10~ps and it is neither a muon nor a neutrino.
} as input to the jet algorithm.
In all cases, jet finding is done in $({\rm rapidity},\phi)$ coordinates and jet calibration is done in $(\ensuremath{\eta^{\rm jet}}\xspace,\ensuremath{\phi^{\rm \ jet}}\xspace)$ coordinates.
The reconstructed jets are additionally required to satisfy several data quality and isolation criteria.
The data quality cuts are each designed to mitigate the impact of specific non-collision backgrounds~\cite{InclJets}.
Reconstructed and particle jets are considered isolated if there is no other reconstructed jet
(or particle jet) within a cone of size $\ensuremath{\Delta R}\xspace=\sqrt{(\Delta\eta)^2+(\Delta\phi)^2}<0.7$ around the jet axis.
Only isolated jets are considered in this study.
The jet vertex fraction (JVF) is calculated for each jet and used to reject
jets originating from pile-up interactions. The JVF is built using information about the origin,
along the direction of the beam, of tracks with $\ensuremath{\Delta R}\xspace<0.4$ ($\ensuremath{\Delta R}\xspace<0.6$) to the jet axis
for $R=0.4$ ($R=0.6$) jets and describes the fraction of the jet's
charged particle $\pt$ associated with the primary vertex~\cite{jesPaper}.
\subsection{Track selection and associating tracks with jets}\label{sec:trackSel}
Tracks are associated with jets by requiring that the track momentum direction (calculated
at the primary vertex) and the jet direction satisfy $\ensuremath{\Delta R}\xspace({\rm jet,track})<0.4$
($\ensuremath{\Delta R}\xspace({\rm jet,track})<0.6$) for $R=0.4$ ($R=0.6$) jets.
Track parameters are evaluated at the point of closest approach to the primary
hard-scattering vertex, which is the vertex with the highest sum of associated track $\pt^2$.
Tracks are required to have $\pt>1$~\GeV, at least one pixel hit and at least six hits
in the silicon strip tracker, as well as transverse (longitudinal) impact parameters with respect
to the hard-scattering vertex $|d_0|<1$~mm ($|z_0\cdot \sin(\theta)|<1$~mm).
The studies in this paper were also performed with a requirement of track $\pt>500$~\MeV. No
significant changes to the results were found. Requiring $\pt>1$~\GeV\ reduces
the sensitivity to pile-up and the underlying event, and this requirement is used for the
remainder of the paper. A ``ghost association''~\cite{ghost} procedure
was also tested in place of $\ensuremath{\Delta R}\xspace$-based matching, and no significant differences are observed.
The jet isolation requirement helps to guarantee the similarity of the ghost association procedure and the $\ensuremath{\Delta R}\xspace$ matching.
\subsection{Photon selection}
Photons with $\pt>25$~\GeV\ are selected with pseudorapidity $|\eta|<2.37$,
excluding the transition region between the barrel and end-cap calorimeters ($1.37<|\eta|<1.52$).
Only the leading photon in the event is considered. The photons are required to satisfy the preselection
and ``tight'' photon cuts described in Ref.~\cite{promptPhot}. An additional isolation cut requiring
less than $5\GeV$ of transverse energy in a cone of size $\ensuremath{\Delta R}\xspace=0.4$ around the photon
is imposed to increase the purity of the sample~\cite{jesPaper}. The photons are additionally
required to be well separated from calorimeter defects and to not be within $\ensuremath{\Delta R}\xspace<0.4$ of a jet arising
from non-collision backgrounds or out-of-time pile-up.
\subsection{Lepton selection}
Isolated electrons and muons are used to select $W+$jet and \antibar{t}\ events.
Electron candidates are formed by matching clusters found in the electromagnetic calorimeter to
tracks reconstructed in the ID in the region $|\eta|< 2.47$ and are required to have
transverse energy $\ET > 25 \GeV$. To ensure good containment of electromagnetic showers in the calorimeter,
the transition region $1.37 <|\eta| < 1.52$ is excluded as for photons.
The electron candidates must pass the ``tight'' selection criteria based on the lateral and transverse
shapes of the clusters described in Ref.~\cite{egammaPaper} but updated for 2011 running conditions.
Reconstructed tracks in the ID and the MS are combined to form muon candidates, which are selected
in the region $|\eta|<2.5$ and are required to have $\pt > 20 \GeV$.
The selection efficiency for electrons and muons in simulated events, as well as their
energy and momentum scale and resolution, are adjusted to reproduce those observed in
$Z\rightarrow\ell\ell$ events in data~\cite{egammaPaper}. To reduce the contamination from
jets identified as leptons, requirements are placed on the total momentum carried by tracks within
$\ensuremath{\Delta R}\xspace=0.3$ of the lepton and on calorimeter energy deposits within $\ensuremath{\Delta R}\xspace=0.2$, excluding the
track and energy of the lepton itself.
For muons, the scalar sum of the $\pt$ of these neighbouring tracks must be less than $2.5 \GeV$, while the
sum of this close-by calorimeter $\ET$ must be less than $4 \GeV$. For electrons, the sum of
calorimeter $\ET$ must be less than $6 \GeV$. Additionally, leptons are required to be
consistent with originating from the primary hard-scattering vertex.
They are required to have $|z_0|<10$~mm, and
the ratio of $d_0$ to its uncertainty ($d_0$ significance) must be smaller than 3.0 for muons
and 10.0 for electrons, due to the wider distribution found in signal electrons caused by
bremsstrahlung.
\subsection{Trigger and event selection}
All events must have a vertex with at least three associated tracks with $\pt>150$~\MeV.
Other event selection requirements are described below.
\subsubsection{Dijet and trijet samples}
The dijet sample is selected using single-jet triggers with various thresholds~\cite{triggerPaper},
which are fully efficient for jets with $\pt>40$~\GeV. Each jet $\pt$ bin is filled exclusively by a single
trigger that is fully efficient for jets in that $\pt$ range, following Ref.~\cite{InclJets}.
The trijet sample uses the same trigger selection as the dijet sample. This guarantees
that studies using the jet with the third highest $\pt$ in each event are not biased by the trigger.
\subsubsection{$\gamma$+jet and $\gamma$+2-jet samples}
The \ensuremath{\gamma\text{+jet}}\xspace\ sample is selected using single-photon triggers. The lowest threshold single-photon trigger
is fully efficient for photons with $\pt>25$~\GeV. For this sample, a back-to-back
requirement for the photon and the leading jet, $\Delta\phi>2.8$, is imposed. An additional
veto on soft radiation is also applied to further reduce background contamination~\cite{jesPaper}:
the uncalibrated $\pt$ of the sub-leading jet is required to be
less than 30\% of the photon $\pt$. Relying on the $\pt$ balance of the photon and jet, each jet $\pt$ bin
is filled exclusively by a single-photon trigger that provides a fully efficient selection.
The same triggers are used in the $\gamma$+2-jet sample in each region of jet $\pt$. Since the
sub-leading jet $\pt$ is lower than that of the leading jet by definition, this selection is also not
biased by jet reconstruction effects.
\subsubsection{$W+$jet sample}
The $W+$jet sample is selected using a single-electron or single-muon trigger.
The event selection, following
Ref.~\cite{Wnote}, requires exactly one charged lepton (electron or muon) and that it matches the
trigger accepting the event, a transverse mass~\footnote{$\mt = \sqrt{\MET \times \ET^\ell \times (1-\cos(\Delta\phi))}$,
where \MET\ is the missing transverse momentum in the event, $\ET^\ell$ is the lepton transverse energy
(transverse momentum for a muon), and $\Delta\phi$ is the angle between the lepton and the \MET\ in the $\phi$ direction.}
$\mt> 40\GeV$, missing transverse momentum $\MET>25\GeV$, and at most two jets (to reject $\antibar{t}$ backgrounds).
The triggers are fully efficient for electrons and muons satisfying the offline $\pt$ requirements.
In events in which two jets are reconstructed, only the jet with the highest $\pt$ is studied.
\subsubsection{$\antibar{t}$ sample}
Top quark pair events in which exactly one of the $\Wboson$ bosons produced by the top quarks decays to
an electron or a muon are selected as described in Ref.~\cite{mtop2011paper}. The event selection requires
that exactly one electron or muon is reconstructed and that it matches the trigger accepting the event. Background
suppression cuts of $\mt>40\GeV$ ($\mt>60\GeV$) and $\MET>25\GeV$
($\MET>20\GeV$) in the electron (muon) channel, and at least four jets with $\ensuremath{\pt^{\rm \ jet}}\xspace> 25\GeV$, $|$JVF$|>0.75$
and $|\ensuremath{\eta^{\rm jet}}\xspace|<2.5$ are also required.
Two of the selected jets must be identified
as arising from a $b$-quark ($b$-tagged) using the
MV1 algorithm, which combines several tracking variables into a multi-variate discriminant,
with the 60\% efficiency working point~\cite{MV1}.
After this selection, the background contamination in the $\antibar{t}$ sample is of the order of 10\% and
consists mainly of events from $W/Z$+jets or single top-quark production. The contribution from
multi-jet background after the requirement of two $b$-tagged jets is about $4\%$~\cite{mtop2011paper}.
The background contamination in the selected data sample has no sizable impact in the studies
performed. The change in the results when including the background in the analysis is small, and the
sample is therefore assumed to be pure $\antibar{t}$.
\section{Jet labelling in Monte Carlo simulation}
\label{sec:jetdef}
One natural definition of the partonic flavour of a jet in a Monte Carlo event is given by matching the
jet to the closest outgoing parton (in $\ensuremath{\Delta R}\xspace$) from the matrix-element calculation, which
represents a fixed-order QCD event record. In generators with $2\rightarrow 2$ matrix elements,
such a matching scheme is clear only for the two leading jets at most.
To simplify the task for analyses using different MC simulations, jets are matched to
the highest-energy parton in the parton shower record within a $\ensuremath{\Delta R}\xspace$ equal to the radius
parameter of the jet algorithm. Using this method, only a small fraction of the jets ($<1$\% around
jet $\pt=50$~\GeV\ and fewer above 100 \GeV) are not assigned a partonic flavour. Studies with {\sc Pythia} 6
and {\sc MadGraph} indicate that jets with significant energy contributions from more than one
distinct parton (e.g. overlap of initial- and final-state radiation) are rare in
the samples used.
The jet isolation requirement restricts the wide-angle QCD radiation of the jet and further guarantees the accuracy
of the labelling based on the parton shower record.
Jets are identified as originating from $c$- and $b$-quarks by requiring one $c$- or $b$-hadron with
$\pt>5$~\GeV\ in the MC record within a $\ensuremath{\Delta R}\xspace$ equal to the radius parameter of the jet.
Jets with two $c$- or $b$-hadrons are identified
as including a gluon splitting to $\antibar{c}$ or $\antibar{b}$. Both classes are considered separately from
quark- and gluon-jets. The labelling of $b$-jets supersedes that of $c$-jets, which itself supersedes the
quark and gluon labelling. In the samples used, other than $\antibar{t}$, the fraction of heavy-flavour jets is relatively small.
The variables used for quark- and gluon-jet discrimination are sufficiently different
for each of these jet types to require an independent treatment.
In MC event generators with matching schemes~\cite{MLM,CKKW,CKKWL}, it is possible to use the outgoing
partons from the matrix-element calculation to label jets. Only jets above the matching
scale can be identified in this manner, and only in exclusively showered events (\text{i.e.}\xspace\ events with the
same number of jets at the matrix-element level and after showering). To avoid the need to tag jets
originating from partons created in the parton shower, the matching scale must be chosen to be
much lower than the minimum $\pt$ of the jets for which the tagger is designed and commissioned.
Labelling of jets based on the highest-energy parton is consistent with labelling based on the
matrix-element calculation for isolated jets in the samples used here. The former is therefore
used in this paper.
For the construction of templates and the examination of data, only ensembles of jets are considered.
The parton record of the MC simulation is not used. Instead, the fractions of quark- and
gluon-jets in each sample are calculated using the matrix-element event record, and only these fractions are
used to describe the average composition of the jet ensemble.
\section{Determination of quark-jet and gluon-jet properties}
\label{sec:mcTemp}
In previous theoretical~\cite{schwartz1} and experimental~\cite{jesPaper} studies, the jet width and
the number of tracks associated with the jet were found to be useful for identifying the partonic origin
of a jet. As discussed in Sec.~\ref{sec:intro}, the larger colour factor associated with a gluon
results in the production of a larger number of particles
and a softer hadron $\pt$ spectrum after the shower. To define the optimal discriminant,
several jet properties are examined for their ability to distinguish the partonic origin of a jet
and for their stability against various experimental effects, including pile-up.
As these jet properties depend on the jet kinematics, the analysis of the properties and the
resulting discriminant are separated into bins of jet $\pt$ and $\eta$. The $\pt$ bin width is dictated
by a combination of the jet resolution and the number of available events in data, and the $\eta$ bins
coarsely follow the detector features.
\subsection{Discriminating variables}
\label{sec:variables}
Useful discriminating variables, such as the number of particles associated with a jet, may be
estimated using either the number of charged-particle tracks in the inner detector or using
the number of topological clusters of energy inside the jet~\cite{jesPaper}.
Although they are
limited to charged particles, and thus miss almost half of the information in a typical
jet, jet properties built from tracks have three practical advantages over calorimeter-based properties.
First, they may include particles that have sufficiently low $\pt$ that they are not measured by
the calorimeter, or which are in the regime where the ID momentum measurement is
more accurate than the energy measurement of the calorimeter. Second, charged
particles bend in the magnetic field of the ID. Additional particles from the
underlying event brought into the jet produce a background in the calorimeter, and
particles that are sufficiently bent are lost to the calorimeter jet. However, both classes of particles
can be correctly assigned using their momenta calculated at the interaction point.
Third, tracks can be easily associated with a specific vertex. This association dramatically
reduces the pile-up dependence of track-based observables. Similar arguments hold in the
calculation of jet shape variables.
The variables surveyed as potential inputs to the quark/gluon tagging discriminant are:
\begin{itemize}
\item Number of reconstructed tracks ($n_{\rm trk}$) in the jet.
\item Calorimeter width:
\begin{equation*}
w = \frac{ \sum_i p_{{\rm T},i} \times \Delta R(i,{\rm jet})}{ \sum_i p_{{\rm T},i} },
\end{equation*}
where the sum runs over the calorimeter energy clusters that are part of the jet.
\item Track width, defined similarly to the calorimeter width but with the sum running over associated tracks.
\item Track-based energy-energy-correlation (EEC) angularity:
\begin{equation*}
{\rm ang_{\rm EEC}}=\frac{ \sum_i \sum_j p_{{\rm T},i} \times p_{{\rm T},j} \times (\Delta R(i,j))^\beta}{ (\sum_i p_{{\rm T},i})^2 },
\end{equation*}
where the index $i$ runs over tracks associated with the jet, $j$ runs over tracks associated with the jet while $j>i$,
and $\beta$ is a tunable parameter~\cite{OriginalEEC,Larkoski2013eya}.
\end{itemize}
The discriminating power (``separation'') of a variable $x$ is calculated
as in Ref.~\cite{2006physics11219P} to investigate the effectiveness of each variable in a quark/gluon tagger in a
sample with equal fractions of quarks and gluons:
\begin{equation*}
s=\frac{1}{2}\int\frac{(p_q(x)-p_g(x))^2}{p_q(x)+p_g(x)}dx = \frac{1}{2}\sum_i\frac{(p_{q,i}-p_{g,i})^2}{p_{q,i}+p_{g,i}},
\label{eq:separation}
\end{equation*}
\noindent where $p_q(x)$ and $p_g(x)$ are the normalised distributions of the variables for quark- and gluon-jets,
and where the second expression applies to histograms, with the sum running over the bins of the histogram.
This definition corresponds to the square of the statistical uncertainty that one would
get in a maximum-likelihood fit when fitting for the fraction of quark- or gluon-jets using the
given variable, divided by the square of the uncertainty in the case of perfect separation.
While this is not a variable that relates easily to quantities of interest for tagging, its interpretation
is independent of the shape of the distributions, allowing for comparisons that are independent
of the tagging efficiency.
Using this definition, Fig.~\ref{fig:someDiscriminants} shows, for different variables, the separation
between quark-jets and gluon-jets as a function of jet $\pt$ for jets built with
the anti-$k_t$ algorithm with $R=0.4$ using the {\sc Pythia} 6 dijet MC simulation.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.45\textwidth]{fig1}
\caption{Separation power provided by different variables between quark- and gluon-jets as
a function of jet $\pt$ in the {\sc Pythia} 6 dijet MC simulation for jets with $|\eta|<0.8$ built
with the anti-$k_t$ algorithm with $R=0.4$. }
\label{fig:someDiscriminants}
\end{center}
\end{figure}
In this simulation, the two most powerful variables are the EEC angularity with $\beta=0.2$ and the
number of tracks associated with the jet.
The jet width built using the associated tracks is the weakest discriminant and the calorimeter-based width is
somewhat stronger, and of comparable power to that of the EEC angularity with $\beta=1.0$.
All track-based variables show excellent stability against pile-up and
significant discrimination power between quark- and gluon-jets.
The dependence of the mean calorimeter width on the number of reconstructed vertices is about five times stronger
than the dependence of any of the variables considered for the final discriminant and at low jet $\pt$ is
up to $\approx 1.5$\% per primary vertex. At high jet $\pt$, the dependence is negligible for all variables.
While it is possible to correct the
inputs or to use a pile-up-dependent selection to allow the use of calorimeter-based
variables without introducing a pile-up dependence in the tagger, such an approach is not pursued in this paper.
Although Fig.~\ref{fig:someDiscriminants} suggests using the charged particle multiplicity and the EEC angularity with
$\beta=0.2$ to build the tagger, a larger linear correlation between these two variables makes this
tagger perform worse at high $\pt$ than the tagger built using the charged particle multiplicity
and the track width. Furthermore, differences between data and MC simulation are
reduced when using the latter tagger. For this reason, track width and $n_{\rm trk}$ are
used to build the discriminant used in the rest of this paper.
The linear correlations between $n_{\rm trk}$ and track width
are at the 15\% level at low $\pt$, increasing to 50\% at high $\pt$. Thus, the variables add
independent information about the properties of the jet.
For EEC angularity with $\beta=0.2$, the linear correlation with $n_{\rm trk}$ is about 75\% with a weak dependence
on $\pt$.
Still, the study of the EEC angularities and the evolution of their tagging performance as a function of
$\beta$ is interesting for reasons discussed in Ref.~\cite{Larkoski2013eya}. Since this discussion
is not relevant for the tagger developed in this paper, it is relegated to \ref{sec:EECresults}.
\subsection{Extraction of pure templates in data}\label{sec:extraction}
To construct a discriminant, the properties of ``pure'' quark- and gluon-jets must be determined.
As these properties depend on the modelling of non-perturbative effects, they are extracted from data
to avoid reliance on MC simulations. The extraction
can be performed using unbiased samples of pure quark- and gluon-jets or, alternatively,
several mixed samples for which the admixture is well known theoretically.
The use of pure samples is explored in detail in Sec.~\ref{sec:disc} as a
validation procedure but is not used to determine the performance of the tagger in data, due
to the limited number of events available and the difficulties in obtaining samples with negligible
gluon and light-quark contaminations. The use of mixed samples is described below in detail, since it is
used to create an operational tagger for data.
Distributions of properties of quark-jets or gluon-jets are extracted using the dijet and
$\gamma+$jet event samples and the fraction of quark- and gluon-jets predicted by
{\sc Pythia} 6 with the {\sc AUET2B} tune.
For each bin $i$ of jet $\eta$, jet $\pt$, and jet property (track width, number of tracks, or the two-dimensional
distribution of the these), a set of linear equations is solved:
\begin{align}
P_i(\eta,\pt) &= f_{q}(\eta,\pt) \times P_{q,i}(\eta,\pt) \nonumber \\
& + f_{g}(\eta,\pt) \times P_{g,i}(\eta,\pt) \nonumber \\
& + f_{c}(\eta,\pt) \times P_{c,i}(\eta,\pt) \nonumber \\
& + f_{b}(\eta,\pt) \times P_{b,i}(\eta,\pt) ,
\label{eq:fractions}
\end{align}
\noindent where $P_i$ is the value of the relevant distribution in bin $i$ of the distribution in the dijet
or $\gamma+$jet sample, $f_{q}$ and $f_{g}$ are the light-quark and gluon fractions
predicted by {\sc Pythia} at a given $\eta$ and $\pt$,
and $P_{q,i}$ and $P_{g,i}$ are the values of the relevant distribution
for quark- and gluon-jets in bin $i$ of the distribution. The fractions
$f_{c}$ and $f_{b}$ for $c$-jets and $b$-jets are relatively small. They are taken from the
MC simulation, together with the corresponding distributions $P_{c}$ and $P_{b}$.
The same is true for the fractions and distributions for $g\rightarrow c\bar{c}$ and $g\rightarrow b\bar{b}$,
not shown in Eq.~\ref{eq:fractions} for brevity.
By using the different fractions of light quarks and gluons in dijet and $\gamma+$jet events in
each $\pt$ and $\eta$ bin, the expected ``pure'' jet sample properties ($P_{q}$ and $P_{g}$)
can be estimated. In these samples, the $b$-jet and $c$-jet fractions are typically below $5-10\%$.
The studies are performed in three bins of $|\eta|$: $|\eta|<0.8$, $0.8<|\eta|<1.2$ and $1.2<|\eta|<2.1$.
An additional term $f_{{\rm fake}, i}(\eta,\pt)\times P_{{\rm fake}, i}(\eta,\pt)$
must be added to the
distributions in the $\gamma+$jet sample to account for events in which the reconstructed
(``fake'') photon arises from a jet with energy deposits mostly within the electromagnetic calorimeter.
The term is estimated from data using a sideband counting technique, developed and implemented in
Refs.~\cite{jesPaper} and~\cite{promptPhot}. The method uses regions defined with
varying levels of photon isolation and photon identification criteria, estimating the number
of background events in the signal region from those in the background regions, after
accounting for signal leakage into the background regions.
Knowledge of $P_x$ and $f_x$ for the dijet and $\gamma+$jet samples allows the
extraction of pure quark- and gluon-jet $n_{\rm trk}$ and track width distributions from the
data. The method can be tested in the MC simulation, comparing the properties
of jets labelled in MC as quark- or gluon-jets and the properties extracted using Eq.~(\ref{eq:fractions})
to demonstrate consistency. Figure~\ref{fig:ntrkWidthClos} (top) shows the mean number of tracks and the mean track width
as a function of the jet $\pt$, separated using either the MC flavour labels
or the extraction procedure in the same MC events for jets with $|\eta|<0.8$.
Differences are observed between the average of the distributions in the dijet and
$\gamma+$jet samples. This biases the extracted distributions for gluon-jets to be more
like the gluon-jet properties in the dijet sample. The same is true for quark-jets and
the $\gamma+$jet sample. The
differences are larger at low $\pt$ and for the track width distributions. The bias demonstrates
a sample dependence, which is included as a systematic uncertainty on
the performance of the discriminant built from these jet properties.
These differences are, however, small compared to the differences between quark- and
gluon-jets, demonstrating the sensitivity of the extraction method. Similar results
are obtained for jets reconstructed with radius parameter $R=0.6$ and in other $|\eta|$ regions.
\begin{figure*}[tp]
\begin{center}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig2a}
}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig2b}
} \\
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig2c}
}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig2d}
}
\caption{ Average (a,c) $n_{\rm trk}$ and (b,d) track width for quark- (solid symbols) and gluon-jets
(open symbols) as a function of reconstructed jet $\pt$ for isolated jets with $|\eta|<0.8$.
Results are shown for distributions obtained using the in-situ extraction method in
{\sc Pythia} 6 simulation (black circles, (a,b)) or data (black circles, (c,d)), as well as for
labelled jets in the dijet sample (triangles) and in the $\ensuremath{\gamma\text{+jet}}\xspace$ sample (squares).
The error bars represent only statistical uncertainties. Isolated jets are reconstructed using the
anti-$k_t$ jet algorithm with radius parameter $R=0.4$. The bottom panels show
the ratio of the results obtained with the in-situ extraction method to the results
in the dijet and $\ensuremath{\gamma\text{+jet}}\xspace$ MC samples. }
\label{fig:ntrkWidthClos}
\end{center}
\end{figure*}
Figure~\ref{fig:ntrkWidthClos} (bottom) shows the same MC simulation points as Fig.~\ref{fig:ntrkWidthClos} (top), but
here the data are used in the extraction. Relatively good agreement is found between data and {\sc Pythia}
{\sc AUET2B} for the track width of gluon-jets and for the number of tracks in quark-jets.
However, the mean number of associated tracks is significantly smaller for gluon-jets in the
data than in the {\sc Pythia} MC simulation. Similarly, the mean
track width is larger in data than in the MC simulation for quark-jets.
Both these differences make gluon-jets and quark-jets more similar, reducing the discrimination
power of these properties in data. Differences between the {\sc Pythia} MC
simulation and the data are also present in some of the other variables originally considered.
These differences translate into non-negligible differences in the corresponding discriminants.
For this reason, a fully data-driven tagger is built.
\subsection{Systematic uncertainties on the extraction procedure}\label{sec:systematicsExtr}
The distributions extracted from data can be used to build a data-driven tagger, and to evaluate its
performance in data.
Uncertainties on the extracted pure quark- and gluon-jet property templates are thus propagated
through as uncertainties on the performance of the tagger. The systematic effects considered can
be classified into four categories:
uncertainties on the input fractions ($f_{x,i}$),
uncertainties on the input shapes ($P_{x,i}$),
uncertainties on the fake photon background,
and sample-dependence effects.
This last category includes, for example, differences in quark-jet properties between samples,
which result in different quark-jet rejection across the various samples. This effect is the one
that causes the inconsistency in the extraction method, illustrated in Fig.~\ref{fig:ntrkWidthClos}.
Sample-dependent effects are included as a systematic uncertainty rather than deriving a separate tagger for
each event selection and MC simulation.
Because jets with different observable properties have different calorimeter response, an
additional uncertainty in the jet energy scale arises from the modelling of the response as a function of the discriminant
in the MC simulation. The resulting uncertainties on the jet energy response after tagging, in addition to the standard
jet energy scale uncertainties, are determined to be below 1\% using a $\ensuremath{\gamma\text{+jet}}\xspace$ $\pt$-balance study following the procedures
described in Ref.~\cite{jesPaper}.
\subsubsection{Input fraction uncertainties}
The fraction of quark- and gluon-jets can change
when going from a leading-order calculation to a next-to-leading-order (NLO) calculation, changing
the renormalisation/factorisation scale, or changing the PDF set.
The first two effects are examined by comparing the {\sc Pythia} and {\sc MadGraph}
calculations, which have different renormalisation/factorisation scales and different ways
of simulating real emissions. Similarly, the potential effect of the real emissions is also probed by
comparing the matrix-element labelling and the highest-energy parton labelling. A 5\% uncertainty
that is anti-correlated between quark- and gluon-jets
is applied to cover the maximum variation seen in these comparisons. This uncertainty
is uncorrelated amongst samples.
The potential mis-modelling of the fraction of quark- and gluon-jets in the MC
simulation due to limitations of the PDFs is estimated using several
PDF sets. The PDF sets use different fitting procedures (MRST, CTEQ and NNPDF sets),
different orders in the perturbation theory expansion (MSTW2008lo for LO, CT10 for NLO)
and different assumptions about the $\alpha_s$ calculation (MRST2007lomod for LO$^*$ and
MRSTMCal for LO$^{**}$). A 5\% uncertainty, anti-correlated between quark- and gluon-jets, conservatively covers the
differences between the various PDF sets. This uncertainty is considered uncorrelated between
the dijet and $\ensuremath{\gamma\text{+jet}}\xspace$ samples because no significant trend is observed between the samples as the
PDF set is changed.
\subsubsection{Heavy-flavour input uncertainties}
The fractions of $b$-jets and $c$-jets are varied by $\pm20$\% in the dijet sample, following Ref.~\cite{HFfraction},
and by $\pm50$\% in the $\gamma+$jet sample to estimate a conservative uncertainty. As
the fractions of $b$-jets and $c$-jets are small, these uncertainties remain sub-leading.
The two input fractions are varied independently. The differences in the results obtained
after the extraction of the pure quark- and gluon-jet properties are added in quadrature
to obtain the total systematic uncertainty from this effect.
Uncertainties on the properties of $b$-jets are determined using a $\antibar{t}$ sample,
described in Sec.~\ref{sec:data-selection}. The purity of this sample is generally better than
95\%. An envelope 10\% uncertainty is included on the $b$-jet properties as a result of
comparisons of $b$-jet properties between data and several MC simulations. The validation is
performed using tagged jets. Differences between tagged and inclusive $b$-jets in
the MC simulation are found to be within the assigned uncertainty.
For $c$-jets, several templates with 10\% increases in the rates of 2-prong, 3-prong, and 4-prong
decays are used to estimate the effect of changes to the $c$-hadron decay.
These different $c$-jet distributions are propagated through the extraction procedure and the
largest difference is used as the systematic uncertainty on the performance of the
tagger due to this effect.
\subsubsection{Fake photon background uncertainty}
Several variations in the background to the $\gamma+$jet sample are considered.
The identification requirements used to define the regions for the background estimation method are
changed, resulting in purity differences of up to 10\% for low-$\pt$ jets. The same procedure
is used to estimate an uncertainty on the jet properties in the fake background. An
uncertainty of up to 4\% covers the changes in the means of the property distributions.
These differences are propagated to the discriminant distribution to obtain a systematic uncertainty
due to the purity estimate. An additional uncertainty covering the full shape correction to
$P_{\rm fake}$ for signal leakage into the background regions of the sideband counting method
is included as well, amounting to less than a 3\% change in the means of the property distributions.
\subsubsection{Sample-dependence uncertainty}
The application to a signal sample of a quark/gluon discriminant derived in a specific set of samples (or sample admixtures) rests
upon the assumption that sample dependence is negligible, or that it can at least be parameterised
as a function of visible properties of the event. One such property is the degree of isolation
of the jet, which requires separate treatment. However, there are other effects, such as colour flow,
that are much harder to constrain using the available data and may lead to a sample-dependence of
jet properties.
Uncertainties on the jet properties are estimated first from differences
between the $\gamma+$jet and dijet samples of the properties of quark- and gluon-jets. These are representative of
the differences observed when comparing several different samples. Events generated
with {\sc Pythia} 6 and {\sc Herwig}++ are also tested for this effect. The envelope of these
variations is used to estimate a systematic uncertainty due to the sample dependence
of the jet properties.
This systematic uncertainty is sensitive to statistical uncertainties in the MC simulation. These statistical
uncertainties
are estimated and used to smooth the $\pt$ dependence of the uncertainty following the
procedure described in Ref.~\cite{JESUpdate}.
The sample dependence is consistently the dominant systematic uncertainty for all jet $\pt$ bins.
The differences between MC labelled samples derive from differences
in observable properties in the dijet and $\ensuremath{\gamma\text{+jet}}\xspace$ samples. It is thus critical to consider these effects when
estimating uncertainties on the tagging efficiency.
The properties of non-isolated jets differ from those of isolated jets, in general.
In both the data and the
MC simulation, isolated jet properties show no significant dependence on the $\ensuremath{\Delta R}\xspace$ to
the nearest reconstructed jet for $\ensuremath{\Delta R}\xspace>0.7$. As the discriminant constructed here uses only jets
satisfying this isolation criterion, no additional uncertainty due to the effect of jet
non-isolation is applied.
An additional uncertainty arises from an incorrect description of the $\pt$-dependence of
the tagging variables for samples with a significantly different jet $\pt$ spectrum from
that of the dijet and $\gamma+$jet samples with which the discriminant was constructed.
This accounts for the differences in bin-to-bin migrations in the various samples.
As this uncertainty is dependent entirely on the sample to which the discriminant is
applied, it is not explicitly included here.
\section{Validation with event-level kinematic cuts}
\label{sec:disc}
The jet property templates extracted in the previous section can be further
validated using high-purity quark- and gluon-jet samples.
Largely following the work in Ref.~\cite{schwartz2}, events are selected using basic kinematic cuts
and event-level selection criteria to study purified samples of quark-jets and gluon-jets.
These event selections are independent of the properties of individual jets and thus do
not bias them. By including several different selections, the importance of colour
flow and other sample-dependent effects can be evaluated using data.
The jets that are not tagged as $b$-jets in the $\antibar{t}$ sample, particularly in the case of
events with exactly four jets, are mostly light-flavour jets. However, because of impurities
introduced by gluon contamination and $\Wboson\rightarrow c\ensuremath{\bar{s}}$ decays, they are not sufficiently
pure to be of use in this study.
\subsection{Validation of gluon-jet properties}\label{sec:gluonJets}
As protons have a large gluon component at low $x$, inclusive low-$\pt$ jet production at the LHC has a
high rate of gluon-jet production. However, the fractions drop rapidly as jet $\pt$ increases.
Particularly at moderate- and high-$|\eta|$, the relative rate of gluon-jet production
exceeds 50\,\% only below 150~\GeV\ in jet $\pt$.
Multi-jet events from QCD contain relatively more gluon radiation than the inclusive jet sample.
The radiation is typically soft, implying that the third-leading jet will
often be a gluon-jet.
A useful kinematic discriminant that can further purify a multi-jet sample, discussed
in Ref.~\cite{schwartz2}, is:
\begin{equation}
\zeta=|\eta_3|-|\eta_1-\eta_2|,
\end{equation}
\noindent where $\eta_i$ is the pseudorapidity of the $i$th leading jet. A selection based on
this variable can provide gluon-jet purity over 90\%, at the price of significantly reduced efficiency.
To evaluate the modelling of gluon-jet properties, events in data with $\zeta<0$ are
compared to those extracted using the template technique described in Sec.~\ref{sec:mcTemp}.
The track multiplicity and jet width are shown in Figs.~\ref{fig:pure_trijet_ntrkTrkWidth1}
and~\ref{fig:pure_trijet_ntrkTrkWidth2}. The
mean values of properties obtained using the purified and (regular) mixed samples generally agree within
statistical and systematic uncertainties. Systematic uncertainties in this figure are calculated as
detailed in Sec.~\ref{sec:systematicsExtr}, and symmetrised around the central value.
\begin{figure*}[tp]
\begin{center}
\subfigure[]{
\label{fig:pure_trijet_ntrkTrkWidth1}
\includegraphics[width=0.45\textwidth]{fig3a}
}
\subfigure[]{
\label{fig:pure_trijet_ntrkTrkWidth2}
\includegraphics[width=0.45\textwidth]{fig3b}
} \\
\subfigure[]{
\label{fig:pure_trijet_ntrkTrkWidth3}
\includegraphics[width=0.45\textwidth]{fig3c}
}
\subfigure[]{
\label{fig:pure_trijet_ntrkTrkWidth4}
\includegraphics[width=0.45\textwidth]{fig3d}
}
\caption{Top, the jet (a) $n_\text{trk}$ and (b) track width as a function of $\pt$ for jets in a
gluon-jet-enriched trijet sample (triangles) compared to gluon-jet extracted templates (circles) for
$|\eta|<0.8$. Bottom, the jet (c) $n_\text{trk}$ and (d) track width as a function of $\pt$ for jets in a quark-jet-enriched $\gamma$+jet
sample (triangles) compared to quark-jet extracted templates (circles) for jets with $|\eta|<0.8$.
Jets are reconstructed with the anti-$k_t$ algorithm with $R=0.4$. The bottom
panels of the figures show the ratios of the results found in the enriched sample to the extracted
results. Error bars on
the points for the enriched sample correspond to statistical uncertainties. The inner
shaded band around the circles and in the ratio represents statistical uncertainties on the
extracted results, while the outer error band represents the combined systematic and statistical
uncertainties.
}
\label{fig:pure_trijet_ntrkTrkWidth}
\end{center}
\end{figure*}
\subsection{Validation of quark-jet properties}\label{sec:quarkEnrichmentGamma}
Events containing photons are widely used as an enriched sample of quark-jets. By selecting
events with photons produced in association with exactly one jet, a sample of quark-jets
that is up to 80\% pure for jets with $\pt>150$~\GeV\ can be constructed. Although the further
enrichment of quark-jets in this sample is difficult, it is possible to obtain higher purities using
events with a photon and two jets~\cite{schwartz2}. If no other selection cuts are applied, these events have a lower
quark-jet fraction than inclusive $\gamma$-jet production. However, a kinematic selection
can help to identify jets seeded by the parton that is most likely to have radiated the photon.
As that parton must have had electric charge, selecting these jets enhances the purity of
quark-jets and rejects gluon-jets.
Following Ref.~\cite{schwartz2}, a variable is defined that allows the kinematic separation of quark-jets and gluon-jets:
\begin{equation*}
\xi=\eta_\text{jet~1}\times\eta_\gamma+\Delta R_{(\text{jet~2},\gamma)},
\end{equation*}
\noindent where $\eta_\gamma$ ($\eta_\text{jet~1}$) is the $\eta$ of the photon (leading jet), and
$\Delta R_{(\text{jet~2},\gamma)}$ gives the difference in $\eta$--$\phi$ space between the sub-leading
jet and the photon. By imposing a requirement on this variable, purities over 90\% can be achieved,
although with a significant loss of events.
To evaluate the modelling of quark-jet properties, events with $\xi<1$ are compared in data
with those extracted using the template technique described in Sec.~\ref{sec:mcTemp}.
The track multiplicity and jet width are shown in Figs.~\ref{fig:pure_trijet_ntrkTrkWidth3}
and~\ref{fig:pure_trijet_ntrkTrkWidth4}.
The two sets of data agree within statistical and systematic uncertainties.
These results also hold in higher $|\eta|$ bins and for jets reconstructed with the anti-$k_t$
algorithm with $R=0.6$.
Additionally, the production of a \Wboson\ boson in association with a jet can be used to provide a
relatively pure sample of quark-jets. A useful variable in constructing the sample is the jet ``charge'', defined as
\begin{equation*}
c_j = \frac{ \sum_i q_i \times \left| \vec{p_i} \cdot \hat{j} \right|^{1/2} } { \sum_i \left| \vec{p_i} \cdot \hat{j} \right|^{1/2} }
\end{equation*}
\noindent where the sums run over all tracks associated with the jet, $\hat{j}$ is a unit three-vector
pointing in the direction of the jet momentum, $\vec{p_i}$ is the track momentum three-vector, and
$q_i$ is the track charge. This variable has been found to be useful in discriminating jets originating
from positively charged quarks from those originating from negatively charged
quarks~\cite{JetCharge1,JetCharge2,ATLASttbar}. The leading contribution to \Wboson\ production
results in a jet with charge opposite to that of the \Wboson\ boson. The main backgrounds are from
gluon-jets, including those in events with jets misidentified as leptons, which should have a
charge distribution
that is approximately Gaussian and centred at zero~\footnote{This is not quite the case, as the
initial state at the LHC is more often positively charged than negatively charged.}.
A pure sample of \Wboson\ events, selected as described in Sec.~\ref{sec:data-selection}, is divided
into events in which the leading jet has a charge with the same sign as the identified lepton (SS) and
those in which the charge is opposite (OS). Templates are then constructed for jet properties in the SS
and OS samples, and the SS sample is used to subtract the gluon-jet contribution from the OS template.
Comparisons between the mean of the OS minus SS distributions in data and MC simulation are
shown in Fig.~\ref{fig:wjetdatamc_ntrk}. The data show reasonable agreement with the
MC simulation, generally within the statistical uncertainties.
The points on these curves disagree at the 10\% level with extracted or purified quark-jet results
shown in previous figures due to a non-closure effect in the method observed at low $\pt$ in the
MC simulation. Results from the $\Wboson+$1-jet MC simulation using generator-based labelling are in agreement
with the quark-jet results from the dijet samples shown in Fig.~\ref{fig:ntrkWidthClos}.
\begin{figure}[htbp]
\begin{center}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig4a}
}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig4b}
}
\caption{The jet (a) $n_\text{trk}$ and (b) track width as a function of $\pt$ for quark-jets in
an OS minus SS \Wboson+jet sample (see text) for $|\eta|<0.8$ in {\sc Pythia} 6 MC simulation and in
data. The panels show the ratio of the results in data to those in MC simulation.}
\label{fig:wjetdatamc_ntrk}
\end{center}
\end{figure}
\section{Light-quark/gluon tagger construction and performance}
\label{sec:perfDetails}
The discriminant for quark- and gluon-jets is based on a simple likelihood ratio
that uses the two-dimensional extracted distributions of $n_{\rm trk}$ and track width for quark-
and gluon-jets:
\begin{equation*}
L=\frac{q}{q+g},
\end{equation*}
\noindent where $q$ ($g$) represents the normalised two-dimensional distribution for quark-jets
(gluon-jets). A selection on $L$ is used
in each bin to discriminate quark- and gluon-jets. This discriminant
is built in bins of jet $\pt$ and $\eta$. The two-dimensional distributions are first smoothed using
a Gaussian kernel and then appropriately rebinned to build the discriminant
distribution in such a way that all bins are populated sufficiently.
The performance of the tagger is determined using the two-dimensional extracted distributions
of $n_{\rm trk}$ and track width in data and those obtained for labelled jets in MC simulations.
Systematic uncertainties on the evaluated performance are estimated using alternative templates
as described in Sec.~\ref{sec:systematicsExtr}.
Table~\ref{tab:perfSummary} summarises this performance for jets with $|\eta|<0.8$.
The efficiencies for gluon-jets and quark-jets are evaluated
only at certain operating points with fixed light-quark efficiency. Statistical uncertainties are
evaluated using pseudoexperiments. Systematic
uncertainties are combined in quadrature and affect both the quark- and
gluon-jet efficiency in data. Large differences between MC simulation and data in
the variables used translate into large scale factors in the gluon-jet efficiency. Practically,
analyses using this tagger would apply the appropriate MC tagger to MC simulation and the data
tagger to data. These scale factors are needed for each MC tagger to create event weights for
the MC simulation, so that the efficiency in the MC simulation matches the measured efficiency
in such analyses. Three representative $\pt$ bins are shown in the table.
\begin{table*}[htbp]
\caption{
Summary of the performance of the quark-jet tagger on quark- and gluon-jets in data and {\sc Pythia}6 MC simulation for jets built with the anti-$k_t$ algorithm with $R=0.4$ and with $|\eta|<0.8$.
The first error corresponds to the statistical uncertainty, while the second corresponds to the systematic uncertainty.
The scale factor is the ratio of data to MC simulation.
\label{tab:perfSummary}}
\begin{center}
\begin{tabular}{|c|l|l|l|l|l|l|}
\hline
& \multicolumn{2}{|c|}{Monte Carlo} & \multicolumn{2}{|c|}{Data} & \multicolumn{2}{|c|}{Scale Factor} \\
\hline
& $\epsilon_{\rm quark}$ & $\epsilon_{\rm gluon}$ & $\epsilon_{\rm quark}$ & $\epsilon_{\rm gluon}$ & SF$_{\rm quark}$ & SF$_{\rm gluon}$ \\
\hline\noalign{\smallskip}
\multirow{4}{*}{\begin{sideways}$\pt=60$--$80 \GeV$\end{sideways}}
& 30\% & 8.4\% & ($30.0\pm0.8^{+3.2}_{-5.3}$)\% & ($11.9\pm0.3^{+7.5}_{-2.9}$)\% & $1.00\pm0.03^{+0.11}_{-0.18}$ & $1.42\pm0.04^{+0.89}_{-0.34}$ \\&&&&&&\\
& 50\% & 21.0\% & ($50.0^{+1.4+4.3}_{-1.3-6.8}$)\% & ($26.6^{+0.8+7.1}_{-0.6-3.9}$)\% & $1.00^{+0.027+0.09}_{-0.026-0.14}$ & $1.27^{+0.04+0.34}_{-0.03-0.19}$ \\&&&&&&\\
& 70\% & 41.5\% & ($70.0^{+1.7+3.9}_{-1.5-11.0}$)\% & ($48.4^{+1.1+4.7}_{-0.9-6.0}$)\% & $1.00^{+0.024+0.06}_{-0.022-0.16}$ & $1.17^{+0.03+0.11}_{-0.02-0.14}$ \\&&&&&&\\
& 90\% & 69.9\% & ($90.0^{+1.5+1.7}_{-1.3-3.3}$)\% & ($80.2^{+1.0+5.6}_{-0.8-2.2}$)\% & $1.00^{+0.02+0.02}_{-0.01-0.04}$ & $1.15^{+0.015+0.08}_{-0.012-0.03}$ \\&&&&&&\\
\hline\noalign{\smallskip}
\multirow{4}{*}{\begin{sideways}\centering$\pt=110$--$160 \GeV$\end{sideways}}
& 30\% & 5.7\% & ($30.0\pm0.6^{+2.8}_{-4.6}$)\% & ($11.6^{+0.6+6.2}_{-0.4-4.6}$)\% & $1.00\pm0.02^{+0.09}_{-0.15}$ & $2.03^{+0.11+1.08}_{-0.08-0.81}$ \\ &&&&&&\\
& 50\% & 13.9\% & ($50.0\pm1.0^{+4.1}_{-6.1}$)\% & ($24.3^{+1.2+7.4}_{-0.8-9.2}$)\% & $1.00\pm0.02^{+0.08}_{-0.12}$ & $1.75^{+0.09+0.53}_{-0.06-0.66}$ \\ &&&&&&\\
& 70\% & 29.7\% & ($70.0^{+1.0+3.9}_{-1.1-8.5}$)\% & ($45.3^{+1.5+4.6}_{-1.1-9.3}$)\% & $1.00^{+0.01+0.06}_{-0.02-0.12}$ & $1.52^{+0.05+0.15}_{-0.04-0.31}$ \\ &&&&&&\\
& 90\% & 64.8\% & ($90.0^{+0.5+2.0}_{-0.6-2.6}$)\% & ($78.1^{+1.0+3.5}_{-0.6-6.0}$)\% & $1.00^{+0.006+0.02}_{-0.007-0.03}$ & $1.21^{+0.02+0.05}_{-0.01-0.09}$ \\ &&&&&&\\
\hline\noalign{\smallskip}
\multirow{4}{*}{\begin{sideways}\centering$\pt=310$--$360 \GeV$\end{sideways}}
& 30\% & 3.9\% & ($30.0^{+5.0+2.1}_{-7.1-4.7}$)\% & ($11^{+5+8}_{-7-4}$)\% & $1.00^{+0.17+0.07}_{-0.24-0.16}$ & $2.8^{+1.4+2.0}_{-1.9-1.1}$ \\&&&&&&\\
& 50\% & 10.3\% & ($50.0^{+8.1+3.0}_{-11.6-8.3}$)\%& ($23^{+10+8}_{-12-9}$)\%& $1.00^{+0.16+0.06}_{-0.23-0.17}$ & $2.2^{+1.0+0.8}_{-1.1-0.9}$ \\&&&&&&\\
& 70\% & 23.5\% & ($70.0^{+7.2+3.1}_{-8.8-7.0}$)\% & ($43^{+8+6}_{-12-10}$)\% & $1.00^{+0.10+0.04}_{-0.13-0.10}$ & $1.81^{+0.35+0.23}_{-0.51-0.42}$ \\&&&&&&\\
& 90\% & 58.9\% & ($90.0^{+5.0+1.8}_{-4.9-3.1}$)\% & ($80^{+6+4}_{-10-7}$)\% & $1.00^{+0.06+0.02}_{-0.05-0.03}$ & $1.37^{+0.10+0.07}_{-0.17-0.11}$ \\&&&&&&\\
\hline
\end{tabular}
\end{center}
\end{table*}
The difference in efficiency between data and MC simulation is particularly large for
the tightest operating point at high $\pt$. It improves for the loosest operating points
and is generally better for the lowest $\pt$ bins.
The efficiencies extracted from data show a much weaker dependence on $\pt$ than is suggested
by {\sc Pythia} 6. No strong dependence on $\eta$ is observed in any sample. The
performance obtained here in {\sc Pythia} 6 compares well with the generator-level studies
presented in Ref.~\cite{schwartz1}. The systematic uncertainties are dominated by the
uncertainty due to the sample dependence.
The efficiencies of the tagger in MC simulation and in data are summarised in Fig.~\ref{fig:perf_summary},
where the performance estimated from labelled jets in dijet MC simulations and extracted data are shown.
Two MC simulation-based taggers were used to produce this figure, one developed using distributions
extracted in {\sc Pythia} 6, which is applied to the {\sc Pythia} 6 samples, and another derived
from {\sc Herwig}++, used for the {\sc Herwig}++ samples.
As expected from Sec.~\ref{sec:extraction}, the data do not agree well with either
{\sc Pythia} 6 or {\sc Herwig}++. Differences between data and {\sc Pythia} 6 are within
systematic uncertainties at low $\pt$, but are more significant at high $\pt$ for those points
for which a large sample is available in the data. The tagger performs worse in {\sc Herwig}++
than on data at low $\pt$ (\ref{fig:perf_summary_a}), but there is fair agreement in its performance
for high $\pt$ jets (\ref{fig:perf_summary_b}).
Comparable results are observed for higher $|\eta|$ ranges, but with larger
statistical uncertainties.
The performance can also be calculated using the relatively pure samples obtained in
trijet and $\gamma+$2-jet events (see Sec.~\ref{sec:disc}).
The efficiencies obtained using purified samples are compared in Fig.~\ref{fig:perf_summary_pur} to those obtained
using the extracted discriminant distribution. The agreement within
systematic uncertainties, particularly in Fig.~\ref{fig:perf_summary_pur_a}, further validates the extraction method.
Some small differences, like those in Fig.~\ref{fig:perf_summary_pur_b}, should be expected from impurities in the
quark and gluon purified samples.
A comparison of performance in jets with radius parameters of $R=0.4$ and $R=0.6$ in data
and MC simulation is shown in Fig.~\ref{fig:perf_summary_rsize}. The performance is
comparable with the two jet sizes.
\begin{figure}[!htp]
\begin{center}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig5a}
\label{fig:perf_summary_a}
}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig5b}
\label{fig:perf_summary_b}
}
\caption{ Gluon-jet efficiency as a function of quark-jet efficiency calculated using jet
properties extracted from data (solid symbols) and from MC-labelled jets
from the dijet {\sc Pythia} 6 (empty squares) and {\sc Herwig}++ (empty diamonds) samples.
Jets with (a) $60<\pt<80$~\GeV\ and (b) $210<\pt<260$~\GeV\ and $|\eta|<0.8$
are reconstructed with the anti-$k_t$ algorithm with $R=0.4$.
The shaded band shows the total systematic uncertainty on the data. The bottom of the plot
shows the ratios of each MC simulation to the data. The error bands on the performance
in the data are drawn around 1.0.
}
\label{fig:perf_summary}
\end{center}
\end{figure}
\begin{figure}[!htp]
\begin{center}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig6a}
\label{fig:perf_summary_pur_a}
}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig6b}
\label{fig:perf_summary_pur_b}
}
\caption{ Gluon-jet efficiency as a function of quark-jet efficiency as calculated using jet properties extracted
from data (solid symbols), purified in data through kinematic cuts (empty diamonds), and extracted from
{\sc Pythia} 6 MC simulation (empty squares).
Jets with (a) $60<\pt<80$~\GeV\ and (b) $210<\pt<260$~\GeV\ and $|\eta|<0.8$ are reconstructed with the anti-$k_t$ algorithm with $R=0.4$.
The shaded band shows the total systematic uncertainty on the data.
The bottom of the plot
shows the ratio of {\sc Pythia} 6 MC simulation or the enriched data samples
to the extracted data. The error bands on the performance
in the data are drawn around 1.0.
}
\label{fig:perf_summary_pur}
\end{center}
\end{figure}
\begin{figure}[!htp]
\begin{center}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig7a}
}
\subfigure[]{
\includegraphics[width=0.45\textwidth]{fig7b}
}
\caption{ Gluon-jet efficiency as a function of quark-jet efficiency as calculated using extracted
jet properties for jets with (a) $60<\pt<80$~\GeV\ and (b) $210<\pt<260$~\GeV\ and $|\eta|<0.8$ reconstructed with
the anti-$k_t$ algorithm with $R=0.4$ (solid symbols) and $R=0.6$ (empty symbols).
The shaded (hashed) band represents the total systematic uncertainty on the $R=0.4$ ($R=0.6$) data points. When
hardly visible, the empty symbols are just behind the solid symbols.
The bottom of the plot
shows the ratio of the performance in data obtained for $R=0.6$ to that for $R=0.4$. Error bands
are drawn around 1.0.
}
\label{fig:perf_summary_rsize}
\end{center}
\end{figure}
\section{Conclusions}
\label{sec:conclusions}
Several variables that are sensitive to differences between quark- and gluon-jets
were studied in various MC simulations and 4.7~fb$^{-1}$ of $\sqrt{s}=7$~\TeV\
$pp$ collision data collected with the ATLAS detector at the LHC during the year 2011.
Two of these variables, chosen to be relatively weakly correlated and stable against pile-up,
were used to build a likelihood-based discriminant to select quark-jets and
reject gluon-jets. Because of non-negligible differences in these variables between
data and MC simulations, a data-driven technique was developed to extract
the discriminant from the data and the MC simulations independently. This technique
exploits significant, $\pt$ dependent differences in the quark- and gluon-jet content
between dijet and $\gamma+$jet samples.
A detailed study of the jet properties reveals that
quark- and gluon-jets look more similar to each other in the data than in the {\sc Pythia}~6
simulation and less similar than in the {\sc Herwig}++ simulation. As a result,
the ability of the tagger to reject gluons at a fixed quark efficiency is up to a factor
of two better in {\sc Pythia} 6 and up to 50\% worse in {\sc Herwig}++ than in data.
Gluon-jet efficiencies in data of $\approx 11\%$ and 80\% are achieved for quark-jet efficiencies of
$\approx 30\%$ and 90\%, respectively. Relative uncertainties of $\approx 5-50\%$ ($\approx 3-20\%$)
were evaluated for the estimate of these gluon-jet (quark-jet) efficiencies,
with the uncertainties increasing for operating points with lower quark-jet efficiency.
These uncertainties are dominated by differences in the properties of quark- and gluon-jets
in the calibration samples (dijet and $\gamma+$jet) and are potentially caused by
effects such as colour flow, which can make radiation around jets different for jets
in different samples, even if they share the same partonic origin. These differences are predicted
to be of different magnitude by the two generators studied and, without further insight, prevent
final-state-dependent taggers to be developed. The differences between the properties in the two
samples are typical of the variations of the properties observed in other samples studied.
The likelihood-based discriminants were
studied independently in kinematically purified gluon-jet and quark-jet samples in data.
Agreement is found within systematic uncertainty between the properties that are used to build the
discriminant for the pure samples and the mixed samples. The same is true for the tagger efficiencies.
Because their properties differ, the same likelihood-ratio discriminant cannot be applied to non-isolated
jets. However, using the methodology described in this paper, a discriminant for non-isolated
jets with typical rejections and efficiencies comparable to those of the isolated-jet
discriminant can be derived.
|
\section{Introduction}
Let
\[
E(x,h;q) = \max_{(a,q)=1} \abs{ \sum_{\substack{x<p\le x+h \\ p\equiv a\mod{q}} } \log p - \frac{h}{\phi(q)} } .
\]
It is believed that $E(x,h;q)\ll_\epsilon x^\epsilon \sqrt{h/q}$ for all $1\le q\le h\le x$, which would imply that each subinterval of $(x,2x]$ of length $>qx^{\epsilon'}$ has its fair share of primes from each reduced arithmetic progression $a\mod q$ (see \cite{Mon;1976,FG;1989} for results and conjectures concerning the case $h=x$). Proving such a result lies well beyond the current technology. However, for several applications it turns out that bounding $E(x,h;q)$ {\it on average} suffices. The case $h=x$, is a rephrasing of the famous Bombieri-Vinogradov theorem: for each fixed $A>0$, there is some $B=B(A)>0$ such that
\[
\sum_{q\le x^{1/2}/(\log x)^B} E(x,x;q) \ll \frac{x}{(\log x)^A} .
\]
Subsequently, various authors focused on obtaining similar results for $h$ small compared to $x$. The first such results were obtained by Jutila \cite{Jut;1969}, Motohashi \cite{Mot;1971}, and Huxley and Iwaniec \cite{HI;1975}. Their bounds were subsequently improved by Perelli, Pintz and Salerno \cite{PPS;1984,PPS;1985} and, finally, by Timofeev \cite{Tim;1987}, who showed that
\eq{bv-short}{
\sum_{q\le Q} E(x,h;q) \ll \frac{h}{(\log x)^A}
}
when $x^{3/5}(\log x)^{2A+129}\le y\le x$ and $Q\le y/(\sqrt{x}(\log x)^{A+64})$, and when $x^{7/12+\epsilon}\le y\le x$ and $Q\le y/x^{11/20+\epsilon}$ with $\epsilon$ fixed and positive.
Results such as the above ones are closely related to what we call {\it zero-density estimates}. Typically, such an estimate is an inequality of the form
\eq{zero-density}{
\sum_{q\le Q} \ssum{\chi\mod q}N(\sigma,T,\chi)\ll (Q^2T)^{c(1-\sigma)} \log^M(QT)
\qquad(Q\ge1,\ T\ge2,\ 1/2\le\sigma\le 1) ,
}
where $c$ and $M$ are some fixed numbers, $N(\sigma,T,\chi)$ denotes the number of zeroes $\rho=\beta+i\gamma$ of the Dirichlet $L$-function $L(s,\chi)$ with $\beta\ge\sigma$ and $|\gamma|\le T$, and the symbol $\ssum{}$ means that we are summing over primitive characters only. The best result of this form we currently know is with $c=12/5+\epsilon$ (and $M=14$ is admissible), a consequence of \cite[Theorem 12.2, eqn. (12.13)]{Mon;1971} when $1/2\le\sigma\le3/4$, of \cite[eqn. (1.1)]{Hux;1974} for $3/4\le\sigma\le 5/6$, and of \cite[Theorem 12.2, eqn. (12.14)]{Mon;1971} when $5/6\le\sigma\le1$. The case $c=2$ and $M=1$ is called the {\it Grand Density Conjecture} \cite[p. 250]{IK;2004}, which, for practical purposes, is often as strong as the Generalized Riemann Hypothesis itself. Proving that \eqref{zero-density} holds for some $c<12/5$ would immediately imply relation \eqref{bv-short} in a wider range of $h$.
\medskip
In the present paper, we study the distribution of primes in short arithmetic progressions too, with the difference that we let the exact location of the interval $(x,x+h]$ vary. The first result of this flavour was shown by Selberg \cite{Sel;1943} when $Q=1$, whose work implies that
\eq{Q=1}{
\int_x^{2x} \abs{\sum_{y<p\le y+h} \log p - h} \mathrm{d} y \ll \frac{hx}{(\log x)^A}
}
for all $A>0$, as long as $h>x^{19/77+\epsilon}$. Huxley's results \cite{Hux;1972} allows one to demonstrate \eqref{Q=1} when $h>x^{1/6+\epsilon}$. Finally, if relation \eqref{zero-density} is true for $Q=1$ and some $c\ge2$, then \eqref{Q=1} holds for $h>x^{1-2/c+\epsilon}$ (see, for example, \cite[Exercise 5]{IK;2004}). Our first result is a generalization of this statement.
\begin{thm}\label{mainthm1}
Assume that relation \eqref{zero-density} holds for some $c\in[2,4]$. Fix $A\ge1$ and $\epsilon\in(0,1/3]$. If $x\ge h\ge1$ and $1\le Q^2\le h/x^{1-2/c+\epsilon}$, then
\[
\int_x^{2x}\sum_{q\le Q} E(y,h;q) \mathrm{d} y \ll_{\epsilon,A} \frac{hx}{(\log x)^A} .
\]
\end{thm}
Theorem \ref{mainthm1} will be proven in Section \ref{mainthm1-proof}. The fact that $h$ is allowed to cover a longer range compared to the case when the interval $(x,x+h]$ is fixed was important in \cite{CDKS2} and \cite{CDKS1}, where primes in progressions were needed to be found in intervals $(x,x+h]$ of length $h\asymp \sqrt{x}$.
Letting $c=12/5+\epsilon$ in Theorem \ref{mainthm1} allows us to take $Q^2=h/x^{1/6+\epsilon'}$. Using a more sophisticated approach based on the frequency of large values of Dirichlet polynomials due to Gallagher-Montgomery and Huxley, it is possible to improve this result when $h>\sqrt{x}$. This is the context of the next theorem, which will be proven in Section \ref{mainthm2-proof}. Note that when $h$ approaches $x$, our result converges towards a weak version of the Bombieri-Vinogradov theorem.
\begin{thm}\label{mainthm2} Fix $A\ge1$ and $\epsilon\in(0,1/3]$. Let $x\ge1$, $h=x^\theta$ with $1/6+2\epsilon\le \theta\le 1$, and $Q\ge 1$ such that $Q^2\le h/x^{\alpha+\epsilon}$, where
\[
\alpha = \begin{cases}
(1-\theta)/3 &\text{if}\ 5/8\le \theta\le 1, \\
1/8 &\text{if}\ 13/24 \le \theta\le 5/8, \\
2/3-\theta &\text{if}\ 1/2\le \theta\le 13/24 ,\\
1/6 &\text{if}\ 1/6+2\epsilon\le \theta\le 1/2.
\end{cases}
\]
Then we have that
\[
\int_x^{2x}\sum_{q\le Q} E(y,h;q) \mathrm{d} y \ll_{\epsilon,A} \frac{hx}{(\log x)^A} .
\]
\end{thm}
\noindent The graph of $\alpha$ as a function of $\theta$ is given below.
\begin{center}
\begin{tikzpicture}
\begin{axis}[xlabel=$\theta$, ylabel=$\alpha$, xmax=1, ymin=0, ymax=0.5,
xtick={0,0.25, 0.5, 0.75, 1}, ytick={0, 0.1, 0.2, 0.5}, xscale=1,yscale=0.5]
\addplot[domain=1/6:1/2]{1/6};
\addplot[domain=1/2:13/24]{2/3-x};
\addplot[domain=13/24:5/8]{1/8};
\addplot[domain=5/8:1]{(1-x)/3};
\addplot[domain=1/2:5/8,dashed]{(1-x)/3};
\end{axis}
\end{tikzpicture}
\end{center}
Finally, we note that using some results due to Li \cite{Li;1997} allows us to take $\alpha=1/15$ in Theorem \ref{mainthm2} if we contend ourselves with only lower bounds on the number of primes in a short arithmetic progressions. A related result was proven by Kumchev \cite{Kum;2002} but for a fixed short interval. It should be noted that when $Q=1$, Jia \cite{Jia;1996} showed that $\alpha=1/20$ is admissible. However, the proof of Jia's =result uses some more specialized results concerning Kloosterman sums that do not have exact analogs when we add a long average over arithmetic progressions as well (though in \cite{HWW;2004} a related result was proven).
\begin{thm}\label{mainthm3} Fix $\epsilon$ and consider $x\ge h\ge 2$ and $1\le Q^2\le h/x^{1/15+\epsilon}$.
Then there is a constant $c=c(\epsilon)>0$ such that, for all $A>0$, we have
\[
\#\left\{ \begin{array}{c} (q,n)\in \mathbb{N}^2 \\ q\le Q,\ n\le x \end{array} :
\sum_{\substack{n<p\le n+h \\ p\equiv a\mod q}}\log p\ge \frac{ch}{\phi(q)}\quad\text{when}\ (a,q)=1\right\}
= Qx + O\left(\frac{Qx}{(\log x)^A}\right) .
\]
\end{thm}
The proof of this result, which will be given in Section \ref{mainthm3-proof}, will be relatively short as we will almost immediately appeal to Li's results and methods from \cite{Li;1997}. We state it and prove it also because of an interesting application it has to a rather distant problem studied in \cite{BPS;2012} and in \cite{CDKS1}. There the quantity of interest was $S(M,K)$, which is defined to be the number of pairs $(m,k)\in \mathbb{N}^2$ with $m\le M$ and $k\le K$ for which there exists an elliptic curve $E$ over $\mathbb{F}_p$ with group of points $E(\mathbb{F}_p)\cong \mathbb{Z}/m\mathbb{Z}\times\mathbb{Z}/mk\mathbb{Z}$. As it was shown by Banks, Shparlinski and Pappalardi in \cite{BPS;2012}, we have that
\[
S(M,K) = \#\{m\le M,\ k\le K:\exists\ p\in(m^2k-2m\sqrt{k}+1,m^2+2m\sqrt{k}+1),\ p\equiv 1\mod{m}\}.
\]
So a straightforward application of Theorem \ref{mainthm3} implies the following result, which is a strengthening of the unconditional part of Theorem 1.5 in \cite{CDKS1}.
\begin{cor} Fix $\epsilon>0$ and $A>0$. If $M\le K^{13/34-\epsilon}$, then we have that
\[
S(M,K) = MK+O_{A,\epsilon}\left(\frac{MK}{(\log K)^A} \right) .
\]
\end{cor}
\subsection*{Acknowledgements} I would like to thank Sandro Bettin for many useful discussions around Lemma \ref{charsums} and the use of a smooth partition of unity, and Kaisa Matom\"aki for various useful suggestions and for providing many useful references. I would also like to thank my coauthors in \cite{CDKS2,CDKS1}, Vorrapan Chandee, Chantal David and Ethan Smith, for their helpful comments as well as for their encouragement.
This work was partially supported by the Natural Sciences and Engineering Research Council of Canada Discovery Grant 435272-2013.
\section{The proof of Theorem \ref{mainthm1}}\label{mainthm1-proof}
Let $x,h, Q, \epsilon$ and $A$ as in the statement of the first part of Theorem \ref{mainthm1}. We may assume that $x$ is large enough in terms of $A$ and $\epsilon$. Throughout the rest of the paper, we set
\[
\mathcal{L} = \log x.
\]
Since $h\ge x^{1-2/c+\epsilon}$ by assumption, relation \eqref{Q=1} and the comments following it imply that it is enough to prove that
\eq{mainthm1-alt1}{
\int_x^{2x} \sum_{q\le Q} E'(y,h;q) \mathrm{d} y \ll_{A,\epsilon} \frac{hx}{\mathcal{L}^A} ,
}
where
\[
E'(x,h;q) := \max_{(a,q)=1} \left| \sum_{\substack{ x < p \le x + h \\ p\equiv a\mod q}} \log p
- \frac{1}{\phi(q)} \sum_{\substack{x<p\le x+h \\(p,q)=1 }}\log p\right| .
\]
(Notice that the condition $(p,q)=1$ in the second sum is trivially satisfied for primes $p>Q$.)
We further reduce this relation to the bound
\eq{mainthm1-alt2}{
\int_x^{4x}\sum_{q\le Q}E''(y,\eta y;q)\mathrm{d} y
\ll_{\epsilon,A} \frac{\eta x^2}{\mathcal{L}^A} ,
}
where
\[
\eta := \frac{h}{x\mathcal{L}^{A+1}} = x^{\theta-1} \mathcal{L}^{-A-1}.
\]
Indeed, covering $(y,y+h]$ by intervals of the form $(y_j,y_{j+1}]$, where $y_j=(1+\eta)^jy$, implies that
\als{
E'(y,h;q) &\le \sum_{1\le (1+\eta)^j\le 1+h/y} E'(y_j,\eta y_j;q)
+ O\left(\frac{\eta x}{\phi(q)}\right) ,
}
by the Brun-Titchmarsch inequality. So
\als{
\int_{x}^{2x} \sum_{q\le Q}
E'(y,h;q) \mathrm{d} y
&\le \sum_{1\le(1+\eta)^j\le 1+h/x} \frac{1}{(1+\eta)^j} \int_{(1+\eta)^jx}^{2(1+\eta)^jx}
\sum_{q\le Q} E'(y,\eta y;q) \mathrm{d} y
+ O(\eta x^2\mathcal{L} ) \\
&\ll \frac{h}{\eta x}\int_x^{4x}\sum_{q\le Q}E'(y,\eta y;q) \mathrm{d} y
+ \eta x^2 \mathcal{L},
}
which implies Theorem \ref{mainthm1} if relation \eqref{mainthm1-alt1} holds. So from now on we focus on proving this relation.
Moreover, we need to remove certain `bad' moduli from our sum. Set
\[
\sigma_0= 1- \frac{c_0\log\mathcal{L}}{\mathcal{L}} ,
\]
where $c_0$ is some large constant to be determined later and let $\mathcal{E}$ be the set of moduli $q\le Q$ which are multiples of integers $d$ modulo which there exists a primitive Dirichlet character $\chi$ such that $N(\sigma_0,x,\chi)\ge1$. Note that any such $d$ must be greater than $D_1:=\mathcal{L}^{4c_0+A+M+3}$, where $M$ is the constant in relation \eqref{zero-density}. This a consequence of the Korobov-Vinogradov zero-free region for $L(s,\chi)$ (see the notes of Chapter 9 in \cite{Mon;1994}) and of Siegel's theorem \cite[p. 126]{Dav;2000}. Therefore
\als{
\int_x^{4x} \sum_{q\in\mathcal{E} } E'(y,\eta y;q) \mathrm{d} y
\ll \sum_{q\in\mathcal{E} }\frac{\eta x^2}{\phi(q)}
& \le \eta x^2\mathcal{L} \sum_{D_1\le d\le Q}\ \sideset{}{^*}\sum_{\chi\mod d} N(\sigma_0,x,\chi)
\sum_{q\le Q,\,d|q}\frac1{q} \\
&\ll \eta x^2\mathcal{L}^2 \sum_{D_1\le d\le Q}\frac1{d}
\ \sideset{}{^*}\sum_{\chi\mod d}N(\sigma_0,x,\chi)
}
by the Brun-Titchmarsch inequality. Splitting the range of $d$ into $O(\mathcal{L})$ intervals of the form $[D,2D]$ and applying relation \eqref{zero-density} with $c\le 4$ to each one of them , we find that
\eq{exceptional}{
\int_x^{4x} \sum_{q\in \mathcal{E} } E'(y,\eta y;q) \mathrm{d} y
\ll \eta x^2\mathcal{L}^{M+2} \max_{D_1 \le D\le 2Q} \frac{(D^2x)^{4(1-\sigma_0)}}{D}
\ll \frac{\eta x^{2+4(1-\sigma_0) }}{ \mathcal{L}^{4c_0+A} } = \frac{\eta x^2}{\mathcal{L}^A} .
}
Therefore, instead of \eqref{mainthm1-alt2}, it suffices to show
\eq{mainthm1-alt2b}{
\int_x^{4x}\sum_{\substack{q\le Q \\ q\notin \mathcal{E} }} E'(y,\eta y;q)\mathrm{d} y
\ll_{\epsilon,A} \frac{\eta x^2}{\mathcal{L}^A} ,
}
Next, let $\Lambda(n)$ be the von Mangoldt function, defined to be $\log p$ if $n=p^k$ for some $k$, and 0 otherwise, and set
\[
E''(x,h;q) = \max_{(a,q)=1} \left| \sum_{\substack{ x < n \le x + h \\ n\equiv a\mod q}} \Lambda(n)
- \frac{1}{\phi(q)} \sum_{\substack{y<n\le y+h \\ (n,q)=1}} \Lambda(n) \right|.
\]
Then we have that
\[
E'(y,\eta y;q) - E''(y,\eta y,q)
\ll \sum_{\substack{y<p^k\le y+\eta y \\ k\ge2}} \log p
\ll (\eta\sqrt{x}+1)\mathcal{L}^2
\]
for every $y\in[x,2x]$. Since $Q^2\le \eta x^{1-\epsilon/2} \le x^{1-\epsilon/2}$, we deduce that
\als{
\int_x^{2x} \sum_{q\le Q} |E'(y,\eta y;q) - E''(y,\eta y;q)| \mathrm{d} y \ll (\eta x^{3/2}Q+xQ)\mathcal{L}^2 \ll_{\epsilon,A} \frac{\eta x^2}{\mathcal{L}^A},
}
thus reducing relation \eqref{mainthm1-alt2b} to showing that
\eq{mainthm1-alt3}{
\int_x^{2x} \sum_{\substack{q\le Q \\ q\notin\mathcal{E} }} E''(y,\eta y;q) \mathrm{d} y \ll_{A,\epsilon} \frac{\eta x^2}{\mathcal{L}^A} .
}
The next step is to switch from arithmetic progressions to sums involving Dirichlet characters. Given an arithmetic function $f:\mathbb{N}\to\mathbb{C}$, we set
\[
S(y,h;f)= \left|\sum_{y<n\le y+h} f(n)\right| .
\]
If $\chi$ is induced by $\chi_1$, then we have that
\eq{(n,q)>1}{
\left| S(y,\eta y;\Lambda\chi) - S(y,\eta y;\Lambda\chi_1) \right|
\le \sum_{ \substack{y< n\le y(1+\eta) \\ (n,q)>1 } } \Lambda(n)
\le \omega(q)\log(2y)\ll \mathcal{L}^2,
}
uniformly in $q\le x$ and $y\le3x$. So for such choices of $q$ and $y$, we have that
\eq{e10}{
E''(y,\eta y;q)
\le \frac1{\phi(q)}\sum_{\substack{\chi\mod q \\ \chi\neq\chi_0}} S(y,\eta y;\chi)
= \frac1{\phi(q)}\sum_{d|q,\,d>1}\ \sideset{}{^*}\sum_{\chi\mod d} S(y,\eta y;\chi)
+ O(\mathcal{L}^2).
}
Therefore, using the inequality $\phi(dm)\ge\phi(d)\phi(m)$, we find that
\als{
\sum_{\substack{q\le Q \\ q\notin\mathcal{E}}} E''(y,\eta y;q)
&\le \sum_{\substack{q\le Q \\ q\notin\mathcal{E}}} \frac1{\phi(q)}\sum_{d|q,\,d>1}
\ssum{\chi\mod d} S(y,\eta y;\Lambda\chi)
+ O(Q\mathcal{L}^2) \\
&\le \sum_{\substack{1<d\le Q\\d\notin\mathcal{E}}} \ssum{\chi\mod d} S(y,\eta y;\Lambda\chi)
\sum_{\substack{q\le Q \\ d|q}} \frac1{\phi(q)}
+ O(Q\mathcal{L}^2) \\
&\ll \mathcal{L} \sum_{\substack{1<d\le Q\\d\notin\mathcal{E}}} \frac{1}{\phi(d)} \ssum{\chi\mod d} S(y,\eta y;\Lambda\chi) + O(Q\mathcal{L}^2) .
}
Hence breaking the interval $(1,Q]$ into $O(\mathcal{L})$ dyadic intervals of the form $(D,2D]$ implies that
\[
\sum_{q\le Q} E''(y,\eta y;q)
\ll \mathcal{L}^3 \max_{1\le D\le Q} \frac{1}{D} \sum_{\substack{D<d\le 2D\\d\notin\mathcal{E}}}
\ssum{\chi\mod d} S(y,\eta y;\Lambda\chi)
+ O(Q\mathcal{L}^2) ,
\]
thus reducing relation \eqref{mainthm1-alt3} to showing that
\eq{mainthm1-alt4}{
\int_x^{4x} \sum_{\substack{D<d\le 2D\\d\notin\mathcal{E}}} \ssum{\chi\mod d} S(y,\eta y;\Lambda\chi) \mathrm{d} y
\ll \frac{D\eta x^2}{\mathcal{L}^{A+3}} \qquad(1\le D\le Q) .
}
In order to prove \eqref{mainthm1-alt4}, we express $S(y,\eta y;\Lambda \chi)$ as a sum over zeroes of $L(s,\chi)$. Let $T_0$ be the unique number of the form $2^j-1$, $j\in\mathbb{N}$, lying in $(x/2,x]$. Applying \cite[p. 118, eqn. (9)]{Dav;2000}, we find that for $q\le x$
\als{
S(y,\eta y;\Lambda \chi)
&\le \left| \sum_{\substack { \rho=\beta+i\gamma\neq0\,:\, L(\rho,\chi)=0 \\ 0\le\beta\le1,\,|\gamma|\le T_0}}
\frac{(1+\eta)^\rho-1}{\rho} \cdot y^\rho \right|
+ O(\mathcal{L}^2) \\
&\le \sum_{1\le2^j\le x/2} \left| \sum_{\substack { \rho=\beta+i\gamma\neq0\,:\, L(\rho,\chi)=0
\\ 0\le\beta\le1,\,2^j\le|\gamma|+1\le2^{j+1}}}
\frac{(1+\eta)^\rho-1}{\rho} \cdot y^\rho \right|
+ O(\mathcal{L}^2)
}
Now, if $L(s,\chi)$ has no zeroes $\rho=\beta+i\gamma$ with $\beta\ge\sigma_0$ and $|\gamma|\le x$, then the functional equation implies that there are no zeroes with $0\le\beta\le1-\sigma_0$ and $|\gamma|\le x$ (except possibly for $\rho=0$, which excluded from our sum). So if we let
\[
\mathcal{Z}(T,\chi)
= \{\rho=\beta+i\gamma:L(\rho,\chi)=0,\,1-\sigma_0\le\beta\le\sigma_0,\,T\le|\gamma|+1\le2T\},
\]
then we find that
\als{
&\int_x^{4x}\sum_{\substack{D<d\le 2D \\ d\notin\mathcal{E} }} \ssum{\chi\mod d}
S(y,\eta y;\Lambda \chi ) \mathrm{d} y \\
&\quad\ll \mathcal{L} \max_{1\le T\le x} \int_x^{4x} \sum_{D<d\le 2D} \ssum{\chi\mod d}
\left| \sum_{\rho\in \mathcal{Z}(T,\chi)} \frac{(1+\eta)^\rho-1}{\rho} \cdot y^\rho \right| \mathrm{d} y
+ D^2x\mathcal{L}^2 .
}
The second error term is $\ll D\eta x^2/\mathcal{L}^{A+3}$, since $D\le Q\le \sqrt{\eta x}/x^{\epsilon/2}$. In order to treat the first error term, we apply the Cauchy-Schwarz inequality. This reduces \eqref{mainthm1-alt4} to showing that
\eq{main 1}{
R(D,T):=\int_x^{4x} \sum_{D<d\le 2D}\ \sideset{}{^*}
\sum_{\chi\mod d} \left| \sum_{\rho\in \mathcal{Z}(T,\chi)} \frac{(1+\eta)^\rho-1}{\rho} \cdot y^\rho \right|^2 \mathrm{d} y
\ll\frac{\eta^2x^3}{\mathcal{L}^{2A+8}}
}
for $1\le D^2\le\eta x^{2/c-\epsilon/2}$ and $1\le T\le x$. Using the identity $|z|^2=z\overline{z}$ to expand the square of the absolute value in~\eqref{main 1} and then integrating over $y\in[x,4x]$, we find that
\als{
R(D,T)
&=\sum_{d\le D}\ \sideset{}{^*}\sum_{\chi\mod d}
\sum_{\rho_1,\rho_2\in \mathcal{Z}(T,\chi)}
\frac{(1+\eta)^{\rho_1}-1}{\rho_1}\frac{(1+\eta)^{\overline{\rho_2}}-1}{\overline{\rho_2}}
\cdot \frac{(4x)^{\rho_1+\overline{\rho_2}+1}
- x^{\rho_1+\overline{\rho_2}+1}}{\rho_1+\overline{\rho_2}+1} \\
&\ll x^3 \mathcal{L}^2 \min\left\{\eta,\frac{1}{T}\right\}^2 \sum_{d\le D}\ \sideset{}{^*}
\sum_{\chi\mod d} \sum_{\rho_1,\rho_2\in \mathcal{Z}(T,\chi)}
\frac{x^{\beta_1+\beta_2-2}}{1+|\gamma_1-\gamma_2|} ,
}
where we have written $\rho_j=\beta_j+i\gamma_j$ for $j\in\{1,2\}$. Since $x^{\beta_1+\beta_2}\le x^{2\beta_1}+x^{2\beta_2}$ and
\[
\sum_{\rho_j\in \mathcal{Z}(T,\chi)}\frac{1}{1+|\gamma_1-\gamma_2|}\ll \mathcal{L}^2 \quad(j\in\{1,2\})
\]
(see for example \cite[p. 98, eqn (1) and (2)]{Dav;2000}), we deduce that
\[
R(D,T)
\ll x^3 \mathcal{L}^4 \min\left\{\eta,\frac{1}{T}\right\}^2
\sum_{d\le D}\ \sideset{}{^*}
\sum_{\chi\mod d} \sum_{\rho\in \mathcal{Z}(T,\chi)} x^{2\beta-2} .
\]
This reduces \eqref{main 1} to proving that
\eq{main 2}{
R'(D,T):=\sum_{d\le D}\ \sideset{}{^*}
\sum_{\chi\mod d} \sum_{\rho\in \mathcal{Z}(T,\chi)} x^{2\beta-2}
\ll \frac{\max\{\eta T,1\}^2}{\mathcal{L}^{2A+12}}.
}
For each Dirichlet character $\chi\mod d$, there are at most $O(T\log(dT))$ zeroes $\rho\in\mathcal{Z}(T,\chi)$ with $\beta\le1/2+1/\mathcal{L}$ by \cite[p. 101, eq. (1)]{Dav;2000}. For the rest of the zeroes, note that $x^{2\beta-2}\ll \mathcal{L}\int_{1/2}^\beta x^{2(\sigma-1)}d\sigma$ and consequently \eqref{zero-density} implies that
\als{
R'(D,T)
&\ll \frac{D^2T\mathcal{L}}{x} + \mathcal{L} \int_{1/2}^{\sigma_0}
\sum_{d\le D}\ \sideset{}{^*}\sum_{\chi\mod d} N(\sigma,T,\chi) \frac{\mathrm{d} \sigma}{x^{2(1-\sigma)}} \\
&\ll \frac{D^2T\mathcal{L}}{x} + \mathcal{L}^{M+1} \int_{1/2}^{\sigma_0}
\left( \frac{ (D^2T)^c } {x^2} \right)^{1-\sigma} \mathrm{d}\sigma.
}
Since $c\in[2,4]$ and $D^2/\eta\le x^{2/c-\epsilon/2}$, we conclude that
\als{
\frac{R'(D,T)}{\max\{1,\eta T\}^2}
&\ll\frac{D^2\mathcal{L}}{\eta x} + \mathcal{L}^{M+1} \int_{1/2}^{\sigma_0}\left( \frac{ (D^2/\eta)^c} {x^2} \right)^{1-\sigma}\mathrm{d}\sigma \\
&\ll\frac{\mathcal{L}}{x^{1-2/c+\epsilon/2}} + \frac{\mathcal{L}^{M+2}}{x^{c\epsilon(1-\sigma_0)/2}}
\ll \frac{1}{\mathcal{L}^{c_0c\epsilon/2-M-2}} .
}
Choosing $c_0=2(2A+M+14)/(c\epsilon)$ then proves \eqref{main 2} thus completing the proof of Theorem \ref{mainthm1}.
\section{Some auxiliary results}\label{aux}
Before we embark on the main part of the proof of Theorems \ref{mainthm2} and \ref{mainthm3}, we state here the main technical tools we will use. The first one is a result due to Montgomery and Gallagher\ \cite[Theorem 7.1]{Mon;1971}.
\begin{lem}\label{large sieve}
Let $\{a_n\}_{n=1}^N$ be a sequence of complex numbers. For $Q\ge1$ and $T\ge1$ we have that
\[
\sum_{q\le Q}\ssum{\chi\mod q}
\int_{-T}^T \abs{ \sum_{n=1}^N\frac{a_n\chi(n)}{n^{it}} }^2\mathrm{d} t
\ll (Q^2T+N) \sum_{n=1}^N|a_n|^2.
\]
\end{lem}
Combining Lemma \ref{large sieve} with a result due to Huxley \cite[Theorem 9.18, p.247]{IK;2004}, we have the following estimate on the frequency of large values of Dirichlets polynomials twisted by Dirichlet characters. Here and for the rest of the paper, given a set
\[
\mathcal{R}\subset\{(t,\chi): t\in\mathbb{R},\ \chi\ \text{is a Dirichlet character}\},
\]
we say that $\mathcal{R}$ is well-spaced if for each $(t,\chi),(t',\chi)\in\mathcal{R}$ with $t\neq t'$, we have that $|t-t'|\ge1$. Moreover, we let $\tau_m$ denote the number of ways to write $n$ as a product of $m$ positive integers.
\begin{lem}\label{large values} Fix $m\in\mathbb{N}$ and $r\ge0$ and let $\{a_n\}_{n=1}^N$ be a sequence of complex numbers such that $|a_n|\le \tau_m(n)(\log n)^r$ for all $n\le N$. For each Dirichlet character $\chi$, we set $A(s,\chi) = \sum_{n=1}^N a_n\chi(n)/n^s$ and we consider a well spaced set
\[
\mathcal{R} \subset \bigcup_{q\le Q}\bigcup_{\substack{\chi\mod q \\ \chi\ \text{primitive}}}
\left\{(t,\chi): t\in\mathbb{R},\ \abs{ A(1/2+it ,\chi )} \ge U \right\} ,
\]
where $U\ge1$, $Q\ge1$ and $T\ge1$ are some parameters. If $H=Q^2T$, then
\[
|\mathcal{R}| \ll_{m,r} \min\left\{ \frac{N+H}{U^2}, \frac{N}{U^2}+\frac{NH}{U^6} \right\} (\log 2N)^{3m^2+6r+18} .
\]
\end{lem}
Finally, we need the following result which allows us to pass from a sum of characters of length $N$ to a shorter sum when $N$ is large enough. Its proof is a standard application of ideas related to the approximate functional equation of $L$-functions (for example, see Section 9.6 in \cite{IK;2004}).
\begin{lem}\label{charsums}
Let $\chi$ be a primitive Dirichlet character modulo $q\in(1,Q]$, $g:[0,+\infty)\to[0,+\infty)$ be a smooth function supported on $[1,4]$, $t\in\mathbb{R}$, $N\ge1$ and $r\in\mathbb{Z}_{\ge0}$. If $|t|\le T$ for some $T\ge2$, and $M=\max\{1,(QT/N)^{1+\delta}\}$ for some fixed $\delta>0$, then
\[
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)(\log n)^r}{n^{1/2+it}}
\ll_{r,\delta} (\log2N)^r
\int_{-\infty}^{\infty}
\left|\sum_{n\le M} \frac{\chi(n)}{n^{1/2+i(u+t)}}\right|
\frac{\mathrm{d} u}{1+u^2} .
\]
\end{lem}
\begin{proof} It suffices to consider the case $r=0$. Indeed, for the general case, note that
\[
(\log n)^r = (\log N+ \log(n/N))^r = \sum_{j=0}^r \binom{r}{j} (\log N)^j (\log(n/N))^{r-j}.
\]
So the general case follows by the case $r=0$ applied with $g\log^{r-j}$, $0\le j\le r$, in place of $g$. Moreover, we may assume that $\delta\le 1/2$.
Set $s_0=1/2+it$ and let $\hat{g}(w)=\int_0^\infty g(u)u^{w-1}\mathrm{d} u$ be the Mellin transform of $g$. Then
\als{
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)}{n^{s_0}}
&= \frac{1}{2\pi i} \int_{\Re(w)=3/4} L(s_0+w,\chi) N^w \hat{g}(w) \mathrm{d} w \\
&= \frac{1}{2\pi i} \int_{\Re(w)=-3/4} L(s_0+w,\chi) N^w \hat{g}(w) \mathrm{d} w ,
}
by Cauchy's theorem, since $\hat{g}$ is entire by our assumption that $g$ is supported on $[1,4]$. We make the change of variable $s=1-w-s_0=\bar{s_0}-w$ so that
\[
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)}{n^{s_0}}
= \frac{N^{\bar{s_0}}}{2\pi i} \int_{\Re(s)=5/4} L(1-s,\chi) N^{-s}\hat{g}(\bar{s_0}-s) \mathrm{d} s .
\]
There is a complex number $\epsilon_\chi$ of modulus 1 such that
\[
L(1-s,\chi) = \frac{2\epsilon_\chi }{q^{1/2}} \left(\frac{q}{2\pi}\right)^{s} \gamma(s,\chi) L(s,\bar{\chi}),
\]
where
\[
\gamma(s,\chi) = \Gamma(s) \cos\left(\frac{\pi(s-a)}{2}\right)
\]
with $a=(1-\chi(-1))/2$. Therefore
\als{
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)}{n^{1/2+it}}
&= \frac{\epsilon_\chi N^{\bar{s_0}} }{q^{1/2} \pi i }
\int_{\Re(s)=5/4} \left(\frac{q}{2\pi N}\right)^s \gamma(s,\chi)
L(s,\bar{\chi}) \hat{g}(\bar{s_0}-s) \mathrm{d} s.
}
We expand the sum $L(s,\bar{\chi})$ and invert the order of integration and summation to find that
\als{
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)}{n^{1/2+it}}
&= \frac{\epsilon_\chi N^{\bar{s_0}} }{q^{1/2}\pi i}
\sum_{n=1}^\infty \bar{\chi}(n)
\int_{\Re(s)=5/4} \left(\frac{q}{2\pi nN}\right)^s \gamma(s,\chi)
\hat{g}(\bar{s_0}-s) \mathrm{d} s.
}
We will show that the terms with $n>M_0:=(QT/N)^{1+\delta}$ contribute very little to the above sum. Indeed, for such an $n$, we shift the line of integration to the line $\Re(s)=A$. If $s=A+iu$, then we have that $\gamma(s,\chi)\ll_{A}1+|u|^{A-1/2}$ by Stirling's formula. Since $\hat{g}(A+iu)\ll_B 1/(1+|u|^B)$, for any $B>0$, we conclude that
\als{
\int_{\Re(s)=5/4} \left(\frac{q}{2\pi nN}\right)^s \gamma(s,\chi)
\hat{g}(\bar{s_0}-s) \mathrm{d} s
&=\int_{\Re(s)=A} \left(\frac{q}{2\pi nN}\right)^s \gamma(s,\chi)
\hat{g}(\bar{s_0}-s) \mathrm{d} s \\
&\ll_A\left(\frac{q}{nN}\right)^A \int_{-\infty}^{\infty} \frac{1+|u|^{A-1/2}}{1+|u-t|^B} \mathrm{d} u
\ll_A \frac{1}{\sqrt{T}} \left(\frac{QT}{nN}\right)^A
}
by taking $B=A+2$ and using our assumptions that $|t|\le T$ and that $q\le Q$. If $A\ge2/\delta$, then we conclude that
\als{
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)}{n^{1/2+it}}
&= \frac{\epsilon_\chi N^{\bar{s_0}} }{q^{1/2}\pi i}
\sum_{n\le M_0} \bar{\chi}(n)
\int_{\Re(s)=5/4} \left(\frac{q}{2\pi nN}\right)^s \gamma(s,\chi)
\hat{g}(\bar{s_0}-s) \mathrm{d} s
+ O_{\delta,A}\left( M^{-\delta A/2} \right) .
}
For the integers $n\le M_0$, we set $s=\bar{s_0}+w$, move the line of integration to the line $\Re(w)=0$ and invert the order of summation and integration to conclude that
\als{
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)}{n^{1/2+it}}
&= \frac{\epsilon_\chi }{(2\pi)^{\bar{s_0}}q^{it}\pi i}
\int_{\Re(w)=0} \left(\frac{q}{2\pi N}\right)^{w}\gamma(\bar{s_0}+w,\chi)
\sum_{n\le M_0} \frac{\bar{\chi}(n)}{n^{\bar{s_0}+w}}
\hat{g}(-w) \mathrm{d} w \\
&\qquad + O_{\delta,A}\left( M^{-\delta A/2} \right) .
}
If $w=-iu$, then $\gamma(\bar{s_0}+w,\chi) \hat{g}(-w) \ll 1/(1+u^2)$. Therefore
\[
\sum_{n=1}^\infty \frac{g(n/N)\chi(n)(\log n)^r}{n^{1/2+it}}
\ll_{\delta,A}
\int_{-\infty}^{\infty}
\left|\sum_{n\le M_0} \frac{\chi(n)}{n^{1/2+i(u+t)}}\right|
\frac{\mathrm{d} u}{1+u^2}
+ \frac{1}{M^{\delta A/2}} .
\]
If $M_0<1$, so that $M=1$, then the lemma follows immediately by the above estimate. If $M_0\ge1$, so that $M=M_0$, then the second term can be absorbed into the main term by taking $A$ large enough: we have that
\[
\int_{-M^2}^{M^2}\left|\sum_{n\le M} \frac{\chi(n)}{n^{1/2+i(u+t)}} \right|^2 du
\gg M^2
\]
by Theorem 9.1 in \cite{IK;2004}. Therefore
\[
\int_{-\infty}^{\infty} \left|\sum_{n\le M} \frac{\chi(n)}{n^{1/2+i(u+t)}} \right| \frac{du}{1+u^2}
\gg \frac{1}{M^{9/2}} \int_{-M^2}^{M^2}\left|\sum_{n\le M} \frac{\chi(n)}{n^{1/2+i(u+t)}} \right|^2 du
\gg \frac{1}{M^{5/2}} ,
\]
which proves the lemma in the case when $M_0\ge1$ too by taking $A=5/\delta$.
\end{proof}
\section{The proof of Theorem \ref{mainthm2}}\label{mainthm2-proof}
Let $A,\epsilon$ and $x,h, Q$ be as in the statement of Theorem \ref{mainthm2}. All implied constants might depend on $\epsilon$ and $A$, as well as on the parameters $k_0,B,C$ and $\delta$, and the function $g$ introduced below.
Arguing as in Section \ref{mainthm1-proof}, we note that is enough to prove that
\[
\int_x^{4x} \sum_{D<d\le 2D} \ssum{\chi\mod d} S(y,\eta y;\Lambda\chi) \mathrm{d} y
\ll \frac{D\eta x^2}{\mathcal{L}^{A+3}} \qquad(1\le D\le Q) ,
\]
where
\[
\eta := \frac{h}{x\mathcal{L}^{A+1}} = x^{\theta-1}\mathcal{L}^{-A-1} .
\]
(Here we don't need to remove the exceptional characters, even though we could do so using relation \eqref{zero-density} with $c=12/5+\epsilon$.) So from now on we fix $D\in[1,Q]$ and we set
\eq{beta-def}{
D^2 = \frac{\eta x}{x^\beta} = x^{\theta-\beta} \mathcal{L}^{-A-1} ,
}
so that $\beta\in[\alpha+\epsilon/2,\theta]$.
Next, we fix $k_0\in\mathbb{N}$ to be chosen later and use Heath-Brown's identity as in \cite[p.1367]{H-B;1982} to replace the von Mangoldt function by certain convolutions. Indeed, we have that
\[
\sum_{y<n\le y+\eta y} \Lambda(n)\chi(n)
= \sum_{k=1}^{k_0} (-1)^{k-1}\binom{k_0}{k}
\sum_{\substack{y<n\le y+\eta y \\ n_{k+1},\dots,n_{2k}\le(4x)^{1/k_0} }} (\log n_1)\mu(n_{k+1})\cdots \mu(n_{2k}) \chi(n_1\cdots n_{2k})
\]
for all $y\le 3x$. We break the range of $n_{k+1},\dots,n_{2k}$ into dyadic intervals $(N_{k+1},2N_{k+1}]$, \dots, $(N_{2k},2N_{2k}]$. For the range of $n_1,\dots,n_{2k}$ we will be more careful and use a smooth partition of unity: there is a smooth function $g:\mathbb{R}_{\ge0}\to\mathbb{R}_{\ge0}$ supported on $[1,4]$ such that
\[
\sum_{j\in\mathbb{Z}} g\left(\frac{x}{2^j}\right) = 1 \qquad(x>0).
\]
Indeed, such a function can be constructed by fixing a smooth function $\tilde{g}:\mathbb{R}_{\ge0}\to\mathbb{R}_{\ge0}$ supported on $[1,2]$ such that $\int_0^\infty \tilde{g}(u) \mathrm{d} u/u=1$, and by setting
\[
g(x) = \int_1^2 \tilde{g}(x/u) \frac{\mathrm{d} u}{u} .
\]
Then we find that $S(y,\eta y;\Lambda\chi)$ can be bounded by $O(\mathcal{L}^{2k_0})$ sums of the form $S(y,\eta y;f\chi)$, where $f=f_1*f_2*\cdots*f_{2k}$ for some $k\in\{1,\dots,k_0\}$ with
\eq{f_j}{
f_j(n) = \begin{cases}
g(n/N_j) \log n &\text{if}\ j=1,\\
g(n/N_j) &\text{if}\ 2\le j\le k,\\
{\bf 1}_{(N_j,2N_j]}(n) \mu(n) &\text{if}\ k+1\le j\le 2k,
\end{cases}
}
and $N_1,\dots,N_{2k}$ being numbers that belong to $[1/4,4x]$ and satisfy the inequalities
\[
N_{k+1},\dots,N_{2k}\le 4x^{1/{k_0}}
\quad\text{and}\quad
N_1\cdots N_{2k} \asymp x.
\]
So it suffices to show that
\eq{mainthm2-alt1}{
\int_x^{4x} \sum_{D<d\le 2D} \ssum{\chi\mod d} S(y,\eta y;f\chi) \mathrm{d} y
\ll \frac{D\eta x^2}{\mathcal{L}^{A+2k_0+3}} .
}
In order to detect the condition $y<n\le y+\eta y$ in $S(y,\eta y;f\chi)$, we use Perron's formula: if we set
\[
F_j(s,\chi) = \sum_{n=1}^\infty\frac{f_j(n) \chi(n) }{n^s} \quad(1\le j\le 2k)
\]
and
\[
F(s,\chi) = F_1(s,\chi)\cdots F_{2k}(s,\chi) = \sum_{n=1}^\infty \frac{f(n)\chi(n)}{n^s},
\]
and we fix some $T_0\in(x/2,x]$, then using the lemma in \cite[p. 105]{Dav;2000} we find that
\eq{e70}{
\sum_{n\le z}f(n)\chi(n)
= \frac1{2\pi i} \int\limits_{\sub{\Re(s) = 1/2\\|\Im(s)|\le T_0}} F(s,\chi) \frac{z^s}{s} \mathrm{d} s
+ O(x^{\epsilon/10}) \qquad(1\le z\le 5x) .
}
Applying this estimate for $z=y$ and $z=y+\eta y$ with $y\in[x,4x]$, we find that
\al{
\int_x^{4x} \sum_{d\le D}\ \ssum{\chi\mod d} S(y,\eta;f\chi) \mathrm{d} y
&= \frac1{2\pi} \sum_{d\le D}\ \ssum{\chi\mod d}
\int_x^{4x} \abs{ \int\limits_{\sub{\Re(s)=1/2 \\ |\Im(s)| \le T_0 }} F(s,\chi)
\frac{(1+\eta)^s-1}{s}\cdot y^s \mathrm{d} s } \nonumber \\
&\quad + O\left(D^2x^{1+\epsilon/10}\right). \label{e80}
}
Dividing the range of integration into $O(\mathcal{L})$ subsets of the form $\{s=1/2+it:T-1\le |t| \le 2T-1\}$, $T=2^m\ge1$, and choosing $T_0$ as the unique number of the form $2^m-1$ belonging to $(x/2,x]$, we find that \eqref{mainthm2-alt1} is reduced to showing that
\eq{mainthm2-alt2}{
\sum_{d\le D}\ \ssum{\chi\mod d}
\int_x^{4x} \abs{\ \int\limits_{\sub{\Re(s)=1/2 \\ T\le|\Im(s)|+1\le2T }} F(s,\chi)
\frac{(1+\eta)^s-1}{s}\cdot y^s \mathrm{d} s }
\ll \frac{D\eta x^2}{\mathcal{L}^{A+2k_0+4}}
}
for all $T\in[1,x/2]$.
We continue by dividing the range of integration according to the size of the Dirichlet polynomials $F_j(s,\chi)$. To this end, we fix some numbers $U_1,\dots,U_{2k}$ with $1\le U_1\ll\sqrt{N_1}\log N_1$ and $1\le U_j\ll \sqrt{N_j}$ for $j\in\{2,3,\dots,2k\}$, and we set
\[
\mathcal{P}(\chi,T,\bs U) = \{t\in\mathbb{R}: T\le|t|+1\le2T,\ U_j\le |F_j(1/2+t,\chi)|+1\le 2U_j\ (1\le j\le 2k)\} .
\]
Then relation \eqref{mainthm2-alt2} is reduced to showing that
\eq{mainthm2-alt3}{
\sum_{D<d\le 2D}\ \ssum{\chi\mod d}
\int_x^{4x} \abs{\,\int\limits_{\sub{s=1/2+it \\ t\in \mathcal{P}(\chi,T,\bs U)} } F(s,\chi)
\frac{(1+\eta)^s-1}{s}\cdot y^s \mathrm{d} s }
\ll \frac{D\eta x^2}{\mathcal{L}^{A+4k_0+4}} ,
}
for all $U_1,\dots,U_{2k}$ as above. From now on we fix such a choice of $U_1,\dots,U_{2k}$, and we set
\[
U=U_1\cdots U_{2k} .
\]
Observe that $|F(1/2+it,\chi)|\asymp U$ for $t\in \mathcal{P}(\chi,T,\bs U)$. We fix two large enough constants $B$ and $C$ to be chosen later and we claim that we may assume that
\eq{U}{
U\le \min\{\mathcal{L}^B\sqrt{x}/D,\sqrt{x}/\mathcal{L}^C \} .
}
First, we show that we may assume that $U\le \mathcal{L}^B\sqrt{x}/D$. Indeed, if $U > \mathcal{L}^B\sqrt{x}/D$, then
\[
\int\limits_{\sub{s=1/2+it \\ t\in \mathcal{P}(\chi,T,\bs U)} } F(s,\chi) \frac{(1+\eta)^s-1}{s}\cdot y^s\mathrm{d} s
\ll \frac{\min\{\eta,1/T\}\sqrt{x}}{\sqrt{x}\mathcal{L}^B/D}\int_{-2T}^{2T} |F(1/2+it, \chi)|^2\mathrm{d} t.
\]
Consequently, Lemma\ \ref{large sieve} implies that
\als{
\sum_{D<d\le 2D}&\ \ssum{\chi\mod d}
\int_x^{4x} \abs{ \,\int\limits_{\sub{s=1/2+it \\ t\in \mathcal{P}(\chi,T,\bs U)}}
F(s,\chi)\frac{(1+\eta)^s-1}{s}\cdot y^s \mathrm{d} s } \\
&\ll \frac{Dx\min\{\eta,1/T\}}{\mathcal{L}^B}\sum_{d\le D}\ \sideset{}{^*}
\sum_{\chi\mod d}\int_{-2T}^{2T} |F(1/2+it,\chi)|^2\mathrm{d} t \\
&\ll \frac{Dx\min\{\eta,1/T \}}{\mathcal{L}^B}(D^2T+x) \mathcal{L}^{4k^2+2}
\ll \frac{Dx}{\mathcal{L}^{B-4k^2-2}}(D^2+\eta x)
\ll \frac{D\eta x^2}{\mathcal{L}^{B-4k^2-2}},
}
which is admissible provided that $B\ge A+8k_0^2+6$, a condition we assume from now on. So we only need to consider the case when $U\le \mathcal{L}^B\sqrt{x}/D$.
Finally, we prove that we may restrict our attention to the case when $U\le \sqrt{x}/\mathcal{L}^C$. If $D>\mathcal{L}^{B+C}$, this is implied by our assumption that $U\le \mathcal{L}^B\sqrt{x}/D$. So we assume that $D\le \mathcal{L}^{B+C}$. We fix a small positive constant $\delta$ to be chosen later and we set
\[
\mathcal{J}=\{1\le j\le 2k: N_j>x^{\delta^2}\}
\quad\text{and}\quad J=|\mathcal{J}|.
\]
As long as $\delta^2<1/(2k)$ (which we shall assume), we have that $J\ge1$. If $j\in\mathcal{J}$ and $\chi$ is a primitive character modulo some $d\in(D,2D]$, then we have that $|F_j(1/2+it,\chi)|\ll \sqrt{N_j}/\mathcal{L}^{C+1}$ for all $t\in[-x,x]$. Showing this inequality is routine: we start by using Perron's formula to express $F_j(1/2+it,\chi)$ in terms of $L'(s,\chi), L(s,\chi)$ or $(1/L)(s,\chi)$, according to whether $j=1$, $1<j\le k$ or $k<j\le2k$, respectively. Then we shift the contour to the left and bound $L(s,\chi),L'(s,\chi)$ or $(1/L)(s,\chi)$ in the neighbourhood of the line $\Re(s)=1$ using exponential sum estimates due to Vinogradov. (Note that when $\Im(s)$ is small and $k<j\le 2k$, we need to make use of our assumption that $D\le\mathcal{L}^{B+C}$ and to apply Siegel's theorem.) Since $|F_j(1/2+it,\chi)| \ll \sqrt{N_j}/\mathcal{L}^{C+1}$ for $t\in[-x,x]$, we may assume that $U_j\ll \sqrt{N_j}/\mathcal{L}^{C+1}$ for all $j\in\mathcal{J}$, which in turn implies that $U=\prod_{j=1}^{2k}\ll \mathcal{L}^{-C-1}\prod_{j=1}^{2k} \sqrt{N_j}\asymp \sqrt{x}/\mathcal{L}^{C+1}$. So if $x$ is large enough, then $U\le \sqrt{x}/\mathcal{L}^C$, as claimed.
Hence from now on we assume that $U$ satisfies relation \eqref{U}. Fix for the moment a character $\chi$ and consider the integral
\[
I=I(\chi,T,\bs U) = \int_{x}^{3x} \abs{\ \int\limits_{\sub{s=1/2+it \\ t\in \mathcal{P}(\chi,T,\bs U)} }
F(s,\chi) \frac{(1+\eta)^s-1}{s} \cdot y^s \mathrm{d} s } \mathrm{d} y.
\]
Employing the Cauchy-Schwarz inequality, we find that
\als{
I^2
&\le 2x\int_{x}^{3x} \abs{\ \int\limits_{\sub{s=1/2+it \\ t\in \mathcal{P}(\chi,T,\bs U)} }
F(s,\chi) \frac{(1+\eta)^s-1}{s} \cdot y^s \mathrm{d} s }^2 \mathrm{d} y \\
&= 2x\int_{x}^{3x} \int\limits_{\sub{s_1=1/2+it_1 \\ t_1\in \mathcal{P}(\chi,T,\bs U)} }
\int\limits_{\sub{s_2=1/2+it \\ t_2\in \mathcal{P}(\chi,T,\bs U)} }
F(s_1,\chi)\overline{F(s_2,\chi)} \frac{(1+\eta)^{s_1}-1}{s_1}
\frac{(1+\eta)^{\overline{s_2}}-1}{\overline{s_2}} y^{s_1+\overline{s_2}}\mathrm{d} s_1\mathrm{d} s_2\mathrm{d} y.
}
We first integrate over $y$ and then observe that
\[
\frac{(1+\eta)^{1/2+it}-1}{1/2+it}\ll \min\left\{\eta,\frac{1}{1+|t|} \right\}
\]
and that $|F(s_1,\chi)\overline{F(s_2,\chi)}|\le |F(s_1,\chi)|^2 +|F(s_2,\chi)|^2$. So we deduce that
\als{
I^2 \ll x^3 \min\left\{\eta,\frac{1}{T} \right\}^2 \int_{\mathcal{P}(\chi,T,\bs U)} |F(1/2+it_1,\chi)|^2
\int_{-2T}^{2T} \frac1{1+|t_1-t_2|}\mathrm{d} t_2\mathrm{d} t_1.
}
The inner integral is $\ll\log(2T)\ll\mathcal{L}$, which reduces \eqref{mainthm2-alt3} to showing that
\[
\sum_{D<d\le 2D}\ \ssum{\chi\mod d} \left( \int_{\mathcal{P}(\chi,T,\bs U)} |F(1/2+it,\chi)|^2 \mathrm{d} t\right)^{1/2}
\ll \frac{D\sqrt{x}\max\{1,\eta T\}}{\mathcal{L}^{A+4k_0+9/2}} .
\]
By an application of the Cauchy-Schwarz inequality, we find that it is enough to show that
\eq{mainthm2-alt4}{
S(T,\bs U):= \sum_{D<d\le 2D}\ \ssum{\chi\mod d} \int_{\mathcal{P}(\chi,T,\bs U)} |F(1/2+it,\chi)|^2 \mathrm{d} t
\ll \frac{\max\{1,\eta T\}^2 x}{\mathcal{L}^{2A+8k_0+9}},
}
for all $T\in[1,x]$ and all $U_1,\dots,U_{2k}$ with $U\le \min\{\mathcal{L}^B\sqrt{x}/D,\sqrt{x}/\mathcal{L}^{C}\}$. We note that this relation follows immediately by Lemma \ref{large sieve} if $T>\mathcal{L}^{A+6k_0^2+6}/\eta$. So from now on we will be assuming that $T\le \mathcal{L}^{A+6k_0^2+6}/\eta$. Moreover, we set
\eq{H-def}{
H=D^2T \le \mathcal{L}^{A+6k_0^2+6}D^2/\eta = x^{1-\beta} \mathcal{L}^{A+6k_0^2+6} ,
}
by the definition of $\beta$ by relation \eqref{beta-def}.
In order to show\ \eqref{mainthm2-alt4}, we will take advantage of the special product structure of $F(s,\chi)$, stemming from the fact that $f$ is a convolution. First, we consider the case when there is some $j\in\{1,\dots,k\}$ with $N_j>(DT)^{1/2+\delta}\mathcal{L}^B$. For simplicity, let us assume that $j\ge2$, the argument when $j=1$ being similar. If $\delta_1$ is chosen so that $(1+\delta_1)/(2+\delta_1)=1/2+\delta$ and $M=\max\{1,(4DT/N_j)^{1+\delta_1}\}$, then Lemma \ref{charsums} implies that
\[
F_j(1/2+it,\chi) = \sum_{n=1}^\infty \frac{g(n/N_j)\chi(n)}{n^{1/2+it}}
\ll \int_{-\infty}^{\infty}
\left|\sum_{n\le M } \frac{ \chi(n)} {n^{1/2+i(u+t)}}\right|
\frac{\mathrm{d} u}{1+u^2} .
\]
Together with the Cauchy-Schwarz inequality and Lemma \ref{large sieve}, this implies that
\als{
S(T,\bs U)
&\ll \int_{-\infty}^{\infty} \sum_{D<d\le 2D}\ssum{\chi\mod d}
\int_{-2T}^{2T}
\left|\sum_{n\le M } \frac{\chi(n)}{n^{1/2+i(u+t)}}\right| ^2
\prod_{\substack{1\le \ell\le 2k \\ \ell\neq j}} |F_\ell(1/2+it,\chi)|^2 \frac{ \mathrm{d} t\mathrm{d} u}{1+u^2} \\
&\ll \left(\frac{x+x(DT/N_j)^{1+\delta_1}}{N_j} + H \right) \mathcal{L}^{4k^2+2}
\le \left(\frac{2x}{\mathcal{L}^{B}} + H \right) \mathcal{L}^{4k^2+2} .
}
Since $H\le x^{1-\beta+o(1)}$ by \eqref{H-def}, we conclude that $S(T,\bs U) \ll x/\mathcal{L}^{B-4k^2-2}$, so \eqref{mainthm2-alt4} does hold, provided that $B\ge 2A+12k_0^2+11$.
The above discussion allows to assume that if $j\in\{1,\dots,k\}$, then
\[
N_j\le (DT)^{1/2+\delta}\mathcal{L}^B \le x^{(1/2+\delta)(1-(\theta+\beta)/2)+o(1)}\le x^{1/2-\delta},
\]
provided that $\delta$ is small enough. We suppose that $k_0\ge 3$, so that $N_j\le x^{1/3}$ for all $j\in\{k+1,\dots,2k\}$. Therefore, we see that $J=|\mathcal{J}|\ge 3$.
Next, we reduce \eqref{mainthm2-alt4} to a problem about large values of Dirichlet polynomials. We set
\[
\mathcal{Z}(\chi,T,\bs U) = \left\{n\in\mathbb{Z}: [n,n+1]\cap \mathcal{P}(\chi,T,\bs U) \neq \emptyset \right\} =: \{n_1,\dots,n_r\},
\]
say, with $n_1<n_2<\cdots<n_r$. For each $j\in\{1,\dots,r\}$, we select $t_j\in[n_j,n_{j+1}]\cap \mathcal{P}(\chi,T,\bs U)$, and we set
$\mathcal{R}_m(\chi,T,\bs U)=\{t_j:1\le j\le r,\ j\equiv m\mod{2}\}$ for $m\in\{0,1\}$, then the sets
\[
\mathcal{R}_m(T,\bs U):= \bigcup_{D<d\le 2D} \bigcup_{\substack{\chi\mod d \\ \chi\ \text{primitive} }}
\{(\chi,t): t\in \mathcal{R}_m(\chi,T,\bs U)\} \qquad(m\in\{0,1\})
\]
are well-spaced according to the definition before Lemma \ref{large values}. Finally, we note that
\[
S(T,\bs U) \ll U^2 (|\mathcal{R}_0(T,\bs U)|+|\mathcal{R}_1(T,\bs U)|),
\]
which reduces \eqref{mainthm2-alt4} to showing that
\eq{mainthm2-alt5}{
|\mathcal{R}_m(T,\bs U)| \ll \frac{x}{\mathcal{L}^{2A+8k_0+9} U^2} \quad(m\in\{0,1\}) .
}
We fix $m\in\{0,1\}$ and proceed to the proof of \eqref{mainthm2-alt5}. We distinguish several cases.
\medskip
\underline{\textsc{Case 1}}. Assume that there is a $j\in \mathcal{J}$ such that $U_j>\sqrt{N_j}/\mathcal{L}^B$.
\medskip
\noindent We fix a large enough positive integer $r$ so that $U_j^{2r}\ge H$ and we apply Lemma \ref{large values} with $F_j(s,\chi)^r$ in place of $A(s,\chi)$ to deduce that
\[
|\mathcal{R}_m(T,\bs U)| \ll \frac{N_j^r+H}{U_j^{2r}} \mathcal{L}^{9r^2+18} \ll \mathcal{L}^{4rB+9r^2+18}
\ll \frac{x}{U^2\mathcal{L}^{2A+8k_0+9}}
\]
by \eqref{U}, provided that $C$ is large enough. So \eqref{mainthm2-alt5} does hold in this case.
\medskip
\underline{\textsc{Case 2}}. Suppose that $U_j\le \sqrt{N_j}/\mathcal{L}^B$ for all $j\in \mathcal{J}$ and that there is some $j\in \mathcal{J}$ such that $U_j\le x^{\beta/2}/\mathcal{L}^{B}$.
\medskip
\noindent We apply Lemma \ref{large sieve} with $\prod_{\ell\neq j} F_\ell(s,\chi)$ in place of $A(s,\chi)$ and use relation \eqref{H-def} to deduce that
\als{
|\mathcal{R}_m(T,\bs U)|\ll \frac{x/N_j+H}{(U/U_j)^2} \mathcal{L}^{12k^2+24}
&= \frac{x}{U^2} \left( \frac{U_j^2}{N_j} + \frac{U_j^2H}{x} \right) \mathcal{L}^{12k^2+24} \\
&\ll \frac{x}{U^2} \left( \frac{1}{\mathcal{L}^{2B}} + \frac{(x^\beta/\mathcal{L}^{2B}) x^{1-\beta}\mathcal{L}^{A+6k_0^2+6}}{x} \right)
\mathcal{L}^{12k^2+24} ,
}
which shows \eqref{mainthm2-alt5} by taking $B\ge 3A/2+13k_0^2+20$.
\medskip
\underline{\textsc{Case 3}}. Assume that $U_j\in[x^{\beta/2}/\mathcal{L}^B,\sqrt{N_j}/\mathcal{L}^{2B}]$ for all $j\in \mathcal{J}$ and that $\beta\ge1/(2J)+\epsilon/2$, where $J=|\mathcal{J}|$.
\medskip
\noindent In this case we argue by contradiction: we begin by assuming that $|\mathcal{R}_m(T,\bs U)|\ge x/(U^2\mathcal{L}^{2A+8k_0+9})$.
For every $j\in \mathcal{J}$, we apply Lemma \ref{large values} with $\prod_{\ell\neq j} F_\ell(s,\chi)$ in place of $A(s,\chi)$ to deduce that
\als{
\frac{x}{U^2\mathcal{L}^{2A+8k_0+9}}
\le |\mathcal{R}_m(T,\bs U)|
&\ll \left(\frac{x/N_j}{(U/U_j)^2} + \frac{Hx/N_j}{(U/U_j)^6}\right)\mathcal{L}^{12k^2+24} \\
&\le \frac{x}{U^2\mathcal{L}^{2B-12k^2-24}} + \frac{x^{2-\beta}U_j^6}{U^6N_j} \mathcal{L}^{A+18k_0^2+30} ,
}
by \eqref{H-def} and our assumption on $U_j$. Since $B\ge 3A/2+13k_0^2+20$, this implies that
\[
\frac{x}{U^2\mathcal{L}^{2A+8k_0+9}} \ll \frac{x^{2-\beta}U_j^6}{U^6N_j} \mathcal{L}^{A+18k_0^2+30}
\qquad\implies\qquad
\frac{U^4}{U_j^6} \ll \frac{x^{1-\beta}}{N_j} \mathcal{L}^{3A+26k_0^2+39} .
\]
We multiply the last inequality for all $j\in \mathcal{J}$ to find that
\[
U^{4J-6} \ll x^{J(1-\beta)-1+k\delta^2} \mathcal{L}^{J(3A+26k_0^2+39)} .
\]
We have that $U\ge\prod_{j\in \mathcal{J}}U_j\ge (x^{\beta/2}/\mathcal{L}^{B})^J$. Therefore
\[
(2J-3)J\beta \le J-1-\beta J + (k+1)\delta^2
\qquad\implies\qquad
2J\beta\le 1 + (k+1)\delta^2,
\]
which contradicts our assumption that $\beta\ge1/(2J)+\epsilon/2$ if $\delta$ is small enough in terms of $\epsilon$ and $k$.
Therefore relation \eqref{mainthm2-alt5} holds in this case too.
\medskip
We note that since $J\ge3$, the above discussion shows Theorem \ref{mainthm1} when $\theta\le 1/2$. Indeed, in this case $\beta\ge \alpha+\epsilon/2=1/6+\epsilon/2\ge1/(2J)+\epsilon/2$, so we see that Cases 1-3 above are exhaustive. Hence from on we may assume that $\theta>1/2$. For such a $\theta$, we must have that $\alpha+\theta\ge 2/3$ by the definition of $\alpha$. Since $\beta\ge\alpha+\epsilon/2$, we must have that $\beta+\theta\ge2/3+\epsilon/2$. So the last case we consider is:
\medskip
\underline{\textsc{Case 4}}. Assume that $U_j\in[x^{\beta/2}/\mathcal{L}^B,\sqrt{N_j}/\mathcal{L}^{2B}]$ for all $j\in \mathcal{J}$, that $\beta<1/(2J)+\epsilon/2$, and that $\beta+\theta\ge2/3+\epsilon/2$.
\medskip
\noindent Recall that $N_j\le (DT)^{1/2+\delta}=x^{(1/2+\delta)(1-(\theta+\beta)/2)+o(1)}$ for all $j\in\{1,\dots,k\}$. If $\delta$ is small enough in terms of $\epsilon$, then our assumption that $\beta+\theta\ge2/3+\epsilon/2$ implies that $N_j\le x^{1/3-\epsilon/10}$ when $1\le j\le k$. If we take $k_0=4$, so that $N_j\le x^{1/4}$ when $k+1\le j\le 2k$, then we must have that $J\ge 4$. In particular, $\beta<1/8+\epsilon/2$, whence we deduce that $\alpha<1/8$. The definition of $\alpha$ then implies that $\theta>5/8$, so that $\alpha=(1-\theta)/3$. Recall that $U\le \mathcal{L}^B\sqrt{x}/D$. Since $U^2\ge (x^{\beta}/\mathcal{L}^{2B})^J \ge x^{4\beta}\mathcal{L}^{-8B}$, we conclude that
\[
x^{4\beta}\le \frac{\mathcal{L}^{10B} x}{D^2} = \mathcal{L}^{10B+A+1} x^{1-\theta+\beta}.
\]
Therefore $\beta\le(1-\theta)/3-o(1)$. However, this contradicts the fact that $\beta\ge \alpha+\epsilon/2\ge (1-\theta)/3+\epsilon/2$. This shows that this last case cannot actually occur, thus completing the proof of relation \eqref{mainthm2-alt5} and hence of Theorem \ref{mainthm2}.
\begin{rem}
The above method cannot be improved without some additional input. This can be seen by taking $N_j=x^{1/J}$ when $1\le j\le J$ and $N_j=1$ when $J<j\le 2k_0$, so that $\mathcal{J}=\{1,\dots,J\}$. Also, we assume that $U_j=N_j^{1/4}$, so that when we apply Lemma \ref{large values}, the two different expressions on the right hand side balance. Then the only estimate we can extract from Lemma \ref{large values} is
\[
|\mathcal{R}_m(T,\bs U)|\ll \left(\frac{x^{r/J}}{U^{2r/J}} + \frac{H}{U^{2r/J}}\right)\mathcal{L}^{O(1)} \quad(r\in\mathbb{N}) .
\]
The largest $r$ we can take while still having that $x^{r/J}/U^{2r/J}$ is smaller than $x/U^2$ is $r=J-1$. For this choice, assuming also that $T=1/\eta$, we find that $H/U^{2r/J} = U^{2/J}H/U^2 = U^{2/J}x^{1-\beta}/U^2$. So in order to make this expression smaller than $x/U^2$, we need to assume that $U^2\le x^{J\beta}$. Since $U_j=N_j^{1/4}$, we have that $U^2\asymp x^{1/2}$, so we must have that $\beta\ge1/(2J)$. As it is clear from the proof, what allows us improve upon this estimate when $\theta>5/8$ is the fact that we also know that $U\le\mathcal{L}^B\sqrt{x}/D$.
\end{rem}
\section{Proof of Theorem \ref{mainthm3}}\label{mainthm3-proof}
Fix $\epsilon,A$ and $x,h,Q$ as in the statement of Theorem \ref{mainthm3}. As in the previous section, all implied constants might depend on $\epsilon$ and $A$.
When $h>x/\mathcal{L}^{A+1}$, then Theorem \ref{mainthm3} follows by the Bombieri-Vinogradov theorem. So assume that $h\le x/\mathcal{L}^{A+1}$. Note that we also have that $h\ge x^{1/15+\epsilon}$. Therefore, using Buchstab's identity as in \cite{Li;1997}, we can construct a function $\rho:\mathbb{N}\to\mathbb{R}$ that satisfies the following properties:
\begin{itemize}
\item $\rho\le {\bf 1}_{\mathbb P}$, where ${\bf 1}_{\mathbb P}$ is the indicator function of the set of primes;
\item $\rho$ is supported on integers free of prime factors $<x^{1/100}$;
\item $\rho(n)=O(1)$ for all $n\le 2x$;
\item there is a positive constant $c_1$ such that
\eq{rho}{
\int_x^{2x} \abs{ \sum_{ y<n \le y+h } \rho(n) - \frac{c_1 h}{\log y} } \mathrm{d} y
\ll \frac{hx}{\mathcal{L}^{A+2}}
}
\end{itemize}
We claim that $\rho$ also satisfies the inequality
\eq{mainthm3-alt1}{
\int_x^{2x}\sum_{q\le Q}
\max_{(a,q)=1} \abs{ \sum_{\substack{ y<n \le y+h \\ n\equiv a\mod q} } \rho(n) - \frac{c_1 h}{\phi(q)\log y} } \mathrm{d} y
\ll_{A,\epsilon} \frac{hx}{\mathcal{L}^A} ,
}
for all $A>0$. Before proving this relation, we will demonstrate that it implies Theorem \ref{mainthm3}. Without loss of generality, we may assume that $h\in\mathbb{Z}$. Consider integers $x_1,x_2,Q_1,Q_2$ with $x\le x_1<x_2\le 2x$ and $Q/2\le Q_1<Q_2\le Q$. Then relation \eqref{mainthm3-alt1} immediately implies that
\[
\sum_{x_1\le m < x_2}\sum_{Q_1<q\le Q_2}
\max_{(a,q)=1} \abs{ \sum_{\substack{ m<n \le m+h \\ n\equiv a\mod q} } \rho(n) - \frac{c_1 h}{\phi(q)\log m} }
\ll \frac{hx}{\mathcal{L}^A} .
\]
Therefore, the number of pairs $(q,m)\in\mathbb{N}^2\cap((Q_1,Q_2]\times[x_1,x_2))$ such that
\[
\abs{ \sum_{\substack{ m<n \le m+h \\ n\equiv a\mod q} } \rho(n) - \frac{c_0h}{\phi(q)\log m} }
\ge \frac{c_1h}{2\phi(q)\log y}
\]
is $O(Qx\mathcal{L}^{-A+1})$. Since $\rho\le {\bf 1}_{\mathbb P}$ and $\log n\sim \log m$ for $n\in[m,m+h]\subset[x,3x]$, we deduce that
\als{
\#\left\{ \begin{array}{c} (q,m)\in \mathbb{N}^2 \\ Q_1<q\le Q_2 \\ x_1\le m< x_2 \end{array} :
\sum_{\substack{m<p\le m+h \\ p\equiv a\mod q}} \log p \ge \frac{c_1h}{3\phi(q)}\ \text{when}\ (a,q)=1\right\}
&= (Q_2-Q_1)(x_2-x_1) \\
&\quad + O\left(\frac{Qx}{\mathcal{L}^{A-1}}\right) .
}
Theorem \ref{mainthm3} then follows with $c=c_1/3$ by the above estimate and a dyadic decomposition argument, after replacing $A$ with $A+1$.
So we have reduced our task to showing relation \eqref{mainthm3-alt1}. In view of relation \eqref{rho}, we further note that we may instead show that
\eq{mainthm3-alt2}{
\int_x^{2x}\sum_{q\le Q}
\max_{(a,q)=1} \abs{ \sum_{\substack{ y<n \le y+h \\ n\equiv a\mod q} } \rho(n) - \frac{1}{\phi(q)}
\sum_{ y<n \le y+h } \rho(n) } \mathrm{d} y
\ll \frac{hx}{\mathcal{L}^A} .
}
Hence, arguing as in Section \ref{mainthm1-proof} and using the zero-density estimate \eqref{zero-density} with $c=12/5+\epsilon$, we find that it suffices to prove that
\eq{mainthm3-alt3}{
\int_x^{4x}\sum_{\substack{D<d\le 2D \\ d\notin\mathcal{E}}}\ssum{\chi\mod d}
S(y,\eta y;\rho\chi) \mathrm{d} y
\ll \frac{D\eta x^2}{\mathcal{L}^{A+4}}
}
with $\eta= h/(x\mathcal{L}^{A+1})$ and $\mathcal{E}$ being the set of moduli $q\le Q$ which are multiples of integers $d$ modulo which there exists a primitive Dirichlet character $\chi$ such that $N(\sigma_0,x,\chi)\ge1$, where $\sigma_0= 1- c_0(\log\mathcal{L})/\mathcal{L}$ for some constant $c_0$ that can be taken arbitrarily large. There is only one thing that does not transfer immediately: relation \eqref{(n,q)>1}. Indeed, we need to be more careful when passing from all Dirichlet characters to primitive ones. Note that in order to perform this passage we need to show that
\[
S:= \sum_{q\le Q} \frac{1}{\phi(q)} \sum_{d|q} \ssum{\chi \mod d} \abs{ \sum_{\substack{y<n\le y+\eta y \\ (n,q)>1}} \chi(n)\rho(n) } \ll \frac{\eta y}{\mathcal{L}^A},
\]
for all $y\in[x,3x]$. We write $d=qm$ and note that the the presence of the character $\chi$ in the inner sum implies automatically that the sum is supported on integers $n$ which are coprime to $d$. So the condition $(n,q)>1$ may be replaced by the condition $(n,m)>1$. Splitting also the range of $d$ into $O(\mathcal{L})$ intervals of the form $(D,2D]$, we deduce that
\[
S \ll \mathcal{L}^2 \max_{1/2\le D\le Q} \frac{1}{D} \sum_{m\le Q/D}\frac{1}{m} \sum_{D<d\le 2D}
\ssum{\chi \mod d} \abs{ \sum_{\substack{y<n\le y+\eta y \\ (n,m)>1}} \chi(n)\rho(n) } .
\]
We fix $D\in[1/2,Q]$ and $m\le Q/D$ and note that the Cauchy-Schwarz inequality and the Large Sieve \cite[p. 160, Theorem 4]{Dav;2000} imply that
\als{
\left(\frac{1}{D} \sum_{D<d\le 2D}
\ssum{\chi \mod d} \abs{ \sum_{\substack{y<n\le y+\eta y \\ (n,m)>1}} \chi(n)\rho(n) } \right)^2
&\ll \sum_{D<d\le 2D}
\ssum{\chi \mod d} \abs{ \sum_{\substack{y<n\le y+\eta y \\ (n,m)>1}} \chi(n)\rho(n) }^2 \\
&\ll \eta y\sum_{\substack{y<n\le y+\eta y \\ (n,m)>1}} |\rho(n)|^2 ,
}
since $D\le Q\le \sqrt{\eta y}$ by assumption. By the properties of $\rho$ mentioned above we find that
\als{
\sum_{\substack{y<n\le y+\eta y \\ (n,m)>1}} |\rho(n)|^2
\ll \sum_{\substack{y<n\le y+\eta y \\ p|n\ \Rightarrow\ p>x^{1/100} \\ (n,m)>1}} 1
\le \sum_{\substack{g|m,\, g>1 \\ p|g\ \Rightarrow\ p>x^{1/100} }}
\sum_{\substack{y<n\le y+\eta y \\ (n,m)=g}} 1
&\ll \sum_{\substack{g|m \\ g>x^{1/100}}} \left(\frac{\eta y}{g} + 1\right)
\ll \frac{\tau(m)\eta y}{x^{1/100}}.
}
Consequently,
\[
S\ll \frac{\mathcal{L}^2\eta y}{x^{1/200}} \sum_{m\le Q/D} \frac{\sqrt{\tau(m)}}{m}
\ll \frac{\eta y}{\mathcal{L}^A},
\]
as claimed. So we may indeed focus on proving relation \eqref{mainthm3-alt3}.
Finally, arguing alone the lines of Section \ref{mainthm1-proof} and of Section \ref{mainthm2-proof}, we find that relation \eqref{mainthm3-alt3} can been reduced to showing that
\[
\sum_{\substack{D<d\le 2D \\ d\notin\mathcal{E}}}\ssum{\chi\mod d} \int_{-2T}^{2T}
\abs{\sum_{n\asymp x} \frac{\rho(n)\chi(n)}{n^{1/2+it}} }^2 \mathrm{d} t \ll \frac{x}{\mathcal{L}^{A_1}}
\quad(1\le T\le \eta^{-1}\mathcal{L}^{A_1},\ D\le Q) ,
\]
for some constant $A_1$ that is sufficiently large in terms of $A$. The fact that $\rho$ does satisfy this bound is now a consequence of the methods in \cite{Li;1997}. Indeed, here the `length of the average' $D^2T$ (we are summing over about $D^2$ characters and integrating over an interval of length $T$) satisfies the inequality $D^2T\le x^{14/15+2\epsilon}$, which is precisely the inequality needed for $T$ in \cite{Li;1997}, which is the length of the average there. We can then employ Lemma \ref{large values} in a completely analogous way to the one Hal\'asz's method is used in \cite{Li;1997}. We also need an additional input: that, for all characters $\chi$ we are averaging over, and for all $N\ge x^\delta$, we have
\[
\sum_{N<p\le 2N} \frac{\chi(p)}{p^{1/2+it}} \ll \frac{\sqrt{N}}{\mathcal{L}^{A_2}} ,
\]
where $A_2$ is a sufficiently large constant. This is a consequence of our assumption that $d>D\ge1$ (so $\chi$ is non-principal) and that $d\notin\mathcal{E}$, provided that $c_0$ is large enough in terms of $A_2$. Thus Theorem \ref{mainthm3} follows.
\begin{rem}
With a little more care, we may even assume that we work with primitive characters $\chi$ such that
\eq{improved prime estimate}{
\sum_{N<p\le 2N} \frac{\chi(p)}{p^{1/2+it}} \ll \frac{\sqrt{N}}{\exp\{\mathcal{L}^{1/3-\epsilon}\}}.
}
Let $z=\exp\{\mathcal{L}^{2/3+\epsilon/2}\}$ and let $\mathcal{E}'$ be the set integers $d$ modulo which there exists a primitive Dirichlet character $\chi$ such that $N(1-c/\log z,x,\chi)\ge1$, where $c$ is a small enough constant. Page's theorem \cite[p. 95]{Dav;2000} and the Korobov-Vinogradov zero-free region for Dirichlet $L$-functions (see the notes of Chapter 9 in \cite{Mon;1994}) imply that $\mathcal{E}'$ contains at most one element $\le z$. Then the argument leading to \eqref{exceptional} allows us to restrict our attention to $q\in[1,Q]$ which are not multiples of integers in $\mathcal{E}'$. Moreover, if $\chi$ is a non-principal character modulo such a $q$, then relation \eqref{improved prime estimate} holds.
\end{rem}
\bibliographystyle{alpha}
|
\section{Introduction: The gradient expansion in $\Lambda$CDM}
\label{sec:intro}
Einstein's general relativity provides a coherent, causal framework in
which to describe classical cosmological dynamics.
$\Lambda$CDM cosmology is a remarkably successful model for our
observed universe,
based on a spatially-flat
Friedmann-Lema\^itre-Robertson-Walker (FLRW) spacetime containing
non-relativistic, collisionless (cold) dark matter (CDM) and a
cosmological constant ($\Lambda$).
While the homogeneous and isotropic FLRW background can be
studied analytically (using relativistic or Newtonian theory) the
fully non-linear evolution of the
inhomogeneous matter distribution in
$\Lambda$CDM cosmology
is usually studied using Newtonian N-body simulations.
These are used to make
detailed predictions for comparison against large-scale galaxy
surveys. As the scale and accuracy of these surveys, and hence that
required from numerical simulations, continues to improve, there has
been growing scrutiny of the reliability of results derived from
Newtonian gravity
\citep{Wands:2009ex,Chisari:2011iq,Green:2011wc,Bruni:2013mua}.
In this letter we examine the characteristic signature of
general relativity in the large-scale matter density and hence the
galaxy distribution.
Within $\Lambda$CDM cosmology the distribution of matter is
described by a pressureless fluid.
The relativistic and Newtonian descriptions of the fluid are very
similar if the appropriate variables are used.
The fluid is characterised by its density, $\rho$, and its motion,
represented
by kinematical variables related to its velocity at every point;
$\Theta$ describes the expansion and $\sigma$ its anisotropic deformation or
shear.
We neglect vorticity to be consistent with the predictions of inflationary cosmology at early times.
If we work in terms of the matter density seen by comoving observers,
$\rho \equiv T_{\mu\nu}u^\mu u^\nu$, and the expansion of the matter
4-velocity, $\Theta\equiv\nabla^\mu u_\mu$, then the relativistic and Newtonian
evolution equations are formally exactly the same \citep{Ell71,maartens:2012}.
The continuity equation for the matter density is
\begin{equation}
\label{covariant:cont}
\dot{\rho} + \Theta \rho = 0\,,
\end{equation}
while the Raychaudhuri equation for the expansion is
\begin{equation}
\label{covariant:Ray}
\dot{\Theta} + \frac13 \Theta^2 + 2\sigma^2 + 4 \pi G \rho - \Lambda = 0\,,
\end{equation}
where a dot denotes derivatives with respect to proper time of the comoving observers, corresponding to a Lagrangian time derivative in the Newtonian description.
The difference between Newton and Einstein formalisms becomes evident in
the constraint equations.
At the heart of Newtonian gravity is the Poisson equation
\begin{equation}
\label{poisson}
\nabla^2_{\rm p} \phi = 4\pi G \rho\;,
\end{equation}
where $\nabla^2_{\rm p}$ is the spatial Laplacian in physical coordinates.
This gives a {\em linear} relation between the gravitational field, $\phi$, and the matter density, $\rho$.
In general relativity the density and expansion are related to the intrinsic curvature of the 3-dimensional space orthogonal to $u^\mu$,
denoted by ${}^{(3)}R$.
This is the energy constraint equation
from Einstein's equations
\begin{equation}
\label{energy:constraint}
\frac23 \Theta^2 - 2\sigma^2 + {}^{(3)}R = 16\pi G \rho + 2 \Lambda\,.
\end{equation}
For the homogeneous and isotropic ($\sigma=0$) background, we have $\rho=\bar{\rho}(t)$ and $\Theta=3H(t)$, where $H$ is the Hubble expansion.
The evolution equations \eqref{covariant:cont} and \eqref{covariant:Ray} then become the familiar FLRW equations
\begin{align}
\dot{\bar{\rho}}=&-3H\bar{\rho},\\
\dot{H}=-H^2&-\frac{4\pi G}{3}\bar{\rho} +\frac{\Lambda}{3},
\end{align}
\noindent while the energy constraint \eqref{energy:constraint} reduces to the
Friedmann constraint,
\begin{equation}
\label{friedmann}
H^2=\frac{8\pi G}{3}\bar{\rho} + \frac{\Lambda}{3}\,,
\end{equation}
with ${}^{(3)}R=0$ for a spatially-flat cosmology.
We characterise this background model by the present day value of
the dimensionless density parameter $\Omega_{m}\equiv{8\pi G
\bar{\rho}_0}/{3 H_0^2}$.
Considering inhomogeneities about the FLRW cosmology, we have
\begin{eqnarray}
\Theta(t,x^i) &=& 3H(t) + \theta (t,x^i)\,, \\
\rho (t,x^i) &=& \bar{\rho}(t) \left[ 1+\delta (t,x^i) \right] \,.
\end{eqnarray}
and the inhomogeneous metric can be written in comoving-synchronous coordinates as
\begin{equation}
\label{metric}
ds^2=-dt^2 +a^2(t) e^{2\zeta(t,x^i)}\, \gamma_{kj} dx^k dx^j \,,
\end{equation}
where $a(t)$ is the cosmological scale factor and $\gamma_{ij}(t,x^i)$ has unit determinant.
The specific initial conditions for $\Lambda$CDM cosmology
are set by a period of inflation in the very early universe.
In particular inflation produces an almost scale-invariant distribution for the primordial metric perturbation $\zeta$ in Eq.~(\ref{metric})
\citep{Lyth:2009zz}. This allows us to consider initially small inhomogeneities on large scales, and to perform a gradient expansion (or long-wavelength approximation)
\citep{Lifshitz:1963ps,Tomita:1975kj,Lyth:1984gv,Salopek:1990jq,Deruelle:1994iz,Bruni:2003hm,Rampf:2012pu},
keeping only leading-order terms, i.e., terms at most second order
in spatial gradients of the metric, in particular $\zeta$.
In this approximation we
have\footnote{Equation (\ref{gradient:order}) refers to quantities defined in comoving-synchronous coordinates. In particular it is the comoving matter density contrast, $\delta$, that determines the growth of large-scale structure \citep{Wands:2009ex}.
The perturbed expansion $\theta$ and the shear $\sigma$ are the trace and traceless scalars of the deformation tensor, which is equivalent to the extrinsic curvature in our gauge. This curvature tensor is given by two spatial gradients of the metric and thus both the perturbed expansion and shear are of second order \cite{Tomita:1975kj}.
}
\citep{Bruni:2013qta}
\begin{equation}
\label{gradient:order}
\delta \sim \theta \sim \sigma \sim {}^{(3)}R \sim \nabla^2\,,
\end{equation}
where $\nabla$ is the spatial gradient in comoving coordinates.
We emphasise that $\delta$, $\theta$ and ${}^{(3)}R$ contain {\em all orders} in a conventional perturbative expansion.
They are leading-order quantities only in terms of spatial gradients.
With this proviso $\delta$ and $\theta$ satisfy the simple
evolution equations, from \eqref{covariant:cont} and
\eqref{covariant:Ray},
\begin{eqnarray}
\label{large:continuity}
\dot\delta + \theta &=& {\cal O}(\nabla^4)\,,
\\
\label{large:Ray}
\dot\theta +2H\theta +4\pi G \bar\rho\delta &=& {\cal O}(\nabla^4)\,.
\end{eqnarray}
These quantities are
subject to the energy constraint, from \eqref{energy:constraint},
\begin{equation}
\label{large:constraint}
\frac{{}^{(3)}R}{4} +H\theta = 4\pi G \bar\rho \delta + {\cal O}(\nabla^4) \,.
\end{equation}
Taking the time derivative of this equation and using the evolution
equations for $\theta$ and $\delta$ we generalise the well-known
first-order result that the conformal curvature,
\begin{equation}
\label{rescaledR}
R \equiv {}^{(3)}R a^2\,,
\end{equation}
remains constant in this large-scale limit
\citep{Lukash:1980iv,Lyth:1984gv,Bruni:1992dg}, i.e.\ $R$ is a first
integral of\eqref{large:continuity} and \eqref{large:Ray} .
\section{The relativistic growing-mode from the non-linear curvature}
\label{sec:curvature}
The equations \eqref{large:continuity}, \eqref{large:Ray} and
\eqref{large:constraint} are well known in perturbation theory: they
are the same linear differential equations and constraint that can be
derived at first order in a conventional perturbative expansion in
synchronous-comoving gauge \citep{Bruni:2013qta}, or in a covariant
gauge-invariant fashion for corresponding quantities
\citep{Bruni:1992dg}. Their solution is therefore formally the same
than in first-order perturbation theory. The two independent solutions
of these linear differential equations are a decaying and a growing
mode \citep{Pee80}. Thanks to inflation, the decaying mode is
negligible, so we focus on the the growing-mode solution. In the
large-scale limit we thus have \citep{Bruni:2013qta}
\begin{equation}
\label{large:delta}
\delta = C(x^i) D_+(t) \,, \quad
\theta = -C(x^i) \dot{D}_+(t),
\end{equation}
where the growth factor, $D_+(t)$, is proportional to the
scale factor, $a(t)$, in an Einstein-de Sitter ($\Omega_m=1$)
cosmology \citep{Bernardeau:2001qr}.
The growing-mode amplitude, $C(x^i)$, is related to the conformal
curvature on large scales through the energy constraint equation
\eqref{large:constraint} evaluated at an initial time $t_{IN}$ early in
the matter-dominated era,
\begin{equation}
\label{large:amplitude}
C(x) = \frac{R}{10a_{IN}^2H_{IN}^2 D_{+IN}} \,.
\end{equation}
The growing mode solution for $\delta$ and
$\theta$ on large scales, Eq.\ \eqref{large:delta}, has the same time dependence as the
first-order perturbative solution, thus it may referred to as the linearly growing mode, however we remark again that in our non-linear case $R$ is only conserved at leading order in our gradient expansion\footnote{In a conventional perturbative expansion, for pressureless matter
$R$ is conserved at all scales at first order \citep{Bruni:1992dg}, but at second order contains a time-dependent part that can be neglected at large scales \citep{Bruni:2013qta}.};
$\delta$,
$\theta$ $R$ and $C$ here contain the large-scale part of all perturbative orders.
In single-field, slow-roll inflation, the primordial metric
perturbation, $\zeta(t,x^i)$, is predicted to have an almost Gaussian
distribution \citep{Maldacena:2002vr,Acquaviva:2002ud}.
Crucially, the conformal curvature $R$, Eq.\ \eqref{rescaledR}, is a {\em non-linear}
function of the spatial metric
in Eq.~\eqref{metric}.
Considering only the scalar part of the initial metric
perturbation at leading order on large scales then the spatial metric
can be taken to have a simplified form\footnote{For scalar perturbations the non-Euclidean part of $\gamma_{ij}$ would be of order $\nabla^2$, hence these terms would give contributions to $R$ of order $\nabla^4$.},
$\gamma_{ij}\simeq\delta_{ij}$, and the conformal curvature
$R$ is then a non-linear function of only the
perturbation $\zeta$ in \eref{metric}. With $\gamma_{ij}\simeq\delta_{ij}$ the function $a(t)\exp[\zeta(x)]$ effectively acts as a local scale factor in the so called ``separate universe" picture corresponding to our gradient expansion. $R$ then represents the corresponding local spatial curvature and
takes a beautifully simple and exact form
\citep{Wald:1984rg}
\begin{eqnarray}
\label{conformal:ricci}
R &\simeq& \exp{(-2\zeta)} \left[ -4 \nabla^2 \zeta - 2 \left( \nabla
\zeta \right)^2 \right] \,
\nonumber
\\
&\simeq & - 4 \Big[ \nabla^2\zeta + \frac12 \left( \nabla \zeta \right)^2
-2 \zeta \nabla^2 \zeta - \nonumber \\
&&\,\zeta \left( \nabla \zeta \right)^2 + 2
\zeta^2 \nabla^2 \zeta +\ldots \Big] \,.
\end{eqnarray}
This expression is second-order in spatial gradients, consistent with
\eqref{gradient:order}, but non-linear in terms of the metric perturbation,
$\zeta$.
Consequently, even if $\zeta$ is described by a Gaussian distribution,
its non-linear relation to the curvature $R$ leads to a non-Gaussian
distribution \citep{inprep} for the comoving density contrast, $\delta$, determined
by the amplitude \eqref{large:amplitude} of the growing mode
\eqref{large:delta}.
At first order in a perturbative expansion
we have
from \eref{conformal:ricci}
\begin{equation}
\label{first:R}
R_1 = -4\nabla^2 \zeta_1 \,.
\end{equation}
Substituting this into the constraint equation (\ref{large:constraint}) we recover the
Poisson equation (\ref{poisson}),
where we identify the Newtonian potential in terms of the first-order
Ricci curvature and the inhomogeneous expansion in the
comoving-synchronous gauge
\begin{eqnarray}
\nabla^{2} \phi_1 &=& a^2 \left[ \frac14 {}^{(3)}R_1 + H\theta_1 \right]
\nonumber\\
&=&
- \nabla^{2} \zeta_1 + a^2 H \theta_1 \,.
\end{eqnarray}
Using the full non-linear expression in \eref{conformal:ricci}, we
can write the conformal curvature in terms of $\zeta$ as an infinite series
\begin{eqnarray}
&&R \simeq - 4 \nabla^2\zeta + \nonumber \\
&&\quad\sum_{m = 0}^{\infty}
\frac{(-2)^{m+1}}{m!} \left[ (m+1) \left( \nabla \zeta \right)^2 -4 \zeta \nabla^2 \zeta \right] \zeta^m .
\end{eqnarray}
It is this non-linear curvature which determines the non-linear
amplitude (\ref{large:amplitude}) of the growing mode density
perturbation (\ref{large:delta}).
\section{The non--linear relativistic effect on structure formation}
\label{sec:deltanl}
We wish to relate this density contrast to the observed
distribution of galaxies revealed by astronomical surveys.
Although a full description requires complex, non-linear astrophysics we can assume that
in $\Lambda$CDM cosmology, galaxies form in virialised dark matter halos which are
biased tracers of the underlying matter distribution on large scales \citep{Peacock:1999ye}.
In the simplest model of spherical collapse in Einstein-de Sitter,
written in comoving-synchronous coordinates, there is an exact
parametric solution \citep{Pee80},
\begin{eqnarray}
\delta & = & \frac{9(\psi-\sin\psi)^2}{2(1-\cos\psi)^3} - 1 \,, \\
t & = & \frac{6^{3/2}}{2R^{3/2}} \left( \psi - \sin\psi \right) \,,
\end{eqnarray}
which can be expanded term by term as
\begin{equation}
\delta = C D_+ + \frac{38}{21} (C D_+)^2 + \ldots \,.
\end{equation}
where the linearly growing mode \eqref{large:delta}, with \eqref{large:amplitude}, is given
by $CD_+ = Ra/10$ for Einstein-de Sitter
in both Newtonian theory and
general relativity \citep{Wands:2009ex}. Halos collapse when $\psi=2\pi$ and the linearly
evolved density contrast reaches a critical value
$\delta_*=1.686$. Thus we can predict the number of collapsed halos
(of a given mass)
at a given time in terms of the number of peaks of the
initial growing mode of the comoving density contrast
(smoothed on a given mass scale)
above a critical value \citep{Press:1973iz}.
Going beyond the
spherical collapse, this is the barrier crossing approach, where halos form
where the linearly growing mode exceeds a critical value.
Note that it is the {\em non-linear} amplitude, $C$, of the
{\em linearly evolved} growing mode \eqref{large:delta} which
determines the halo density and this is given by the full non-linear
conformal curvature, $R$ in \eref{conformal:ricci}. In a general
relativistic description of spherical collapse \citep{Wands:2009ex} it
is thus the initial density contrast in the local comoving matter,
$\delta$, that predicts the distribution of halos \citep{Bruni:2011ta}
and as we have seen, this is non-linearly related to the primordial
metric perturbation $\zeta$.
To understand the effect of this non-linearity on structure formation,
we consider a peak-background split \citep{Peacock:1999ye}, where one
decomposes a field into shorter wavelength modes, which generate local
peaks, and much longer
wavelength modes which modulate the number density of peaks\footnote
{Equivalent conclusions can be obtained by studying the distribution of peaks of the metric perturbation, setting $(\nabla\zeta)^2=0$, or by studying the squeezed limits of higher-order correlation functions of the density field \citep{Bruni:2013qta}.
}.
Note that, since we have already made a gradient expansion in the
above (wavenumbers $k<k_{\rm max}$), spatial gradients of our ``shorter wavelength'' modes should
still be small ($k_{\rm split}<k<k_{\rm max}$), and we will now drop completely all gradients of the very
long wavelength modes ($k<k_{\rm split}$).
For simplicity, from now on we shall assume the simplest inflationary scenario where $\zeta$ is Gaussian, focusing on the specific general-relativistic non-Gaussianity introduced by the non-linearity of \eref{conformal:ricci} in the constraint \eqref{large:constraint}.
We can split
\begin{equation}
\label{split}
\zeta \equiv \zeta_\ell + \zeta_s \,,
\end{equation}
where the longer and shorter wavelength modes are
independent for an
initially Gaussian metric perturbation. Substituting \eref{split} into \eref{conformal:ricci}
we obtain
\begin{eqnarray}
&&R \simeq \exp(-2\zeta_\ell) R_s + \nonumber \\
&&\quad 4 \exp(-2\zeta_\ell-2\zeta_s) \nabla\zeta_\ell\nabla\zeta_s + \exp(-2\zeta_s) R_\ell ,
\end{eqnarray}
where
\begin{equation}
\label{short:ricci}
R_s = \exp{(-2\zeta_s)} \left[ -4 \nabla^2 \zeta_s - 2 \left( \nabla \zeta_s \right)^2 \right] \,,
\end{equation}
and similarly for $R_\ell$.
Dropping all spatial gradients of long-wavelength modes,
$\nabla\zeta_\ell$, i.e., taking these modes to define a locally
homogeneous background, we find that the spatial curvature due to
short wavelength modes is modulated such that $R\simeq
\exp(-2\zeta_\ell)R_s$.
This is consistent with the interpretation that the long wavelength
metric perturbation is a rescaling of the local
background scale factor \citep{Maldacena:2002vr,Creminelli:2004yq,Bartolo:2005fp,Creminelli:2011sq,Creminelli:2013mca}
\begin{equation}
\label{local:scale}
a \to a_\ell = \exp(\zeta_\ell) a
\,.
\end{equation}
Hence the local amplitude of the
growing mode of the density
contrast is also modulated (cf. \eqref{large:delta} and
\eqref{large:amplitude})
\begin{equation}
\label{short:delta}
\delta = \exp(-2\zeta_\ell) \delta_s + \mathcal{O}(\nabla\zeta_\ell)\,.
\end{equation}
The non-linear effect of a long-wavelength overdensity,
$\zeta_\ell >0$, suppresses the amplitude of
shorter-wavelength modes since $\zeta_\ell>0$ increases the local
effective scale factor, suppressing spatial curvature and thus the
density contrast.
We can compare \eref{short:delta} with local-type primordial
non-Gaussianity \citep{Wands:2010af} in a Newtonian approach where the
amplitude of the linearly growing mode of the density is determined by
the Newtonian potential
\begin{equation}
\label{local:phi}
\phi = \phi_1 + f_{\rm NL} \left( \phi_1^2 - \langle \phi_1^2 \rangle \right)
+ g_{\rm NL} \phi_1^3 + h_{\rm NL} \left( \phi_1^4 - \langle \phi_1^4 \rangle \right) + \ldots \,.
\end{equation}
If we split the first-order Newtonian potential into longer and shorter
wavelength modes $\phi_1=\phi_\ell+\phi_s$ and drop the gradients of
$\phi_\ell$, we find
\begin{equation}
\label{local:delta}
\delta = \left( 1 + 2f_{\rm NL} \phi_\ell + 3g_{\rm NL} \phi_\ell^2
+ 4h_{\rm NL} \phi_\ell^3 +\ldots \right) \delta_s + \ldots
\end{equation}
The modulation of the amplitude of smaller-scale density fluctuations
$\delta_s$ by the long-wavelength potential, $\phi_\ell$, modifies the
halo density giving rise to a strong modulation of the halo power
spectrum on sufficiently large scales, where $\phi_\ell$ remains
finite even though the long-wavelength density contrast is suppressed,
$\delta\sim\nabla^2$. This leads to a scale-dependent bias in the
distribution of galaxies on large scales
\citep{Dalal:2007cu,Matarrese:2008nc}.
If we impose the same linear relation between $\zeta$ and the
Newtonian potential, $\phi=(3/5)\zeta$,
that is valid for first-order perturbations in the matter-dominated era,
then in single-field, slow-roll inflation $f_{\rm NL}$ and all
higher-order coefficients are suppressed. This results in an
effectively Gaussian distribution for the Newtonian potential and
hence (in Newtonian theory) the density field. However, expanding the
exponential in \eref{short:delta} and comparing term by term with the
equivalent Newtonian expression \eref{local:delta} we can identify
the effective non-Gaussianity on large scales in
general relativity
\begin{equation}
\label{eff:fNL}
f_{\rm NL}^{\rm GR}= -\frac53
\,,\quad
g_{\rm NL}^{\rm GR}= \frac{50}{27}
\,, \quad
h_{\rm NL}^{\rm GR}= -\frac{125}{81}
\,, \cdots
\end{equation}
More generally we find
\begin{equation} \label{gen:fNL}
f_{\rm NL}^{(n)\ {\rm GR}}= \frac{1}{n!} \left( -\frac{10}{3} \right)^{n-1} \,,
\end{equation}
where we write the local expansion (\ref{local:phi}) as
\begin{equation}\label{gen:phiNL}
\phi = \phi_1 + \sum_{n=2}^\infty f_{\rm NL}^{(n)} \left( \phi_1^n - \langle\phi_1^n \rangle \right) \,,
\end{equation}
extending the previous result at second order for $f_{\rm NL}^{\rm
GR}$ \citep{Bartolo:2005xa,Verde:2009hy,Bartolo:2010rw,Hidalgo:2013,Bruni:2013qta,Uggla:2014hva} to higher
orders.
\section{Discussion}
\label{sec:disc}
Traditionally, primordial non-Gaussianity is described in terms of the Newtonian gravitational potential, $\phi$ [for example equation (\ref{gen:phiNL})], linearly related to the density field through the Poisson equation (\ref{poisson}). On the other hand, inflationary predictions are expressed in terms of the primordial metric perturbation, $\zeta$ in Eq.~(\ref{metric}).
Our results, valid in full non-linearity and at large scales,
show how the essential non-linearity of general relativity produces an intrinsic non-Gaussianity in the matter density field and hence the galaxy distribution on large scales, even starting from purely Gaussian primordial metric perturbations, generalising previous results in second-order perturbation theory \citep{Tomita:2005et,Bartolo:2005xa,Verde:2009hy,Bartolo:2010rw,Hwang:2012bi,Hidalgo:2013,Bruni:2013qta,Uggla:2014hva}
We also need a detailed modelling of observational surveys, including all geometrical and relativistic effects, to fully disentangle effects of primordial non-Gaussianity from intrinsic non-linearity in general relativity \citep{Bruni:2011ta,Raccanelli:2013dza,inprep}.
Most studies of GR effects on observations of large-scale structure have been restricted to linear perturbation theory \citep{Yoo:2009au,Yoo:2010ni,Bonvin:2011bg,Challinor:2011bk},
but there have been recent attempts to include non-linear effects, see e.g.\ \citep{Thomas:2014aga,Bertacca:2014dra,Bertacca:2014wga,Yoo:2014sfa,DiDio:2014lka,Jeong:2014ufa}.
Alternative gravity theories may impose different constraints between
the primordial metric perturbation $\zeta$ and the comoving density
contrast $\delta$, and hence could in principle be distinguished by a
different galaxy distribution on large scales.
This could be an interesting approach to testing gravity on
cosmological scales,
complementary to existing work which probes gravity through the growth
of cosmic structure at late times \citep{Zhao:2011te}.
Even within the context of general relativity the constraint equation
(\ref{energy:constraint}) could include additional contributions due
to other fields such as dark energy/quintessence.
Fields which have a negligible effect in the background could still
contribute to the inhomogeneous perturbations, e.g., magnetic fields
or gravitational waves. In particular we have considered only the
growing mode of scalar perturbations at early times. Tensor metric
perturbations are decoupled from scalar density perturbations at first
order, but do contribute to the Ricci curvature at second order, even
in the large scale limit, and hence could contribute to the non-linear
density perturbation \citep{Matarrese:1997ay,Dai:2013kra}, although this is expected to be sub-dominant.
In summary, in this letter we have obtained for the first time the fully non-linear general relativistic
initial distribution of primordial density perturbations in $\Lambda$CDM
on large scales
\begin{equation}
\delta= \frac{ \exp{(-2\zeta)} \left[ -4 \nabla^2 \zeta - 2 \left( \nabla
\zeta \right)^2 \right] }{10a_{IN}^2H_{IN}^2 D_{+IN}} D_+(t)\,,
\end{equation}
an expression including {\em all perturbative orders}.
This fully non-linear relation between $\delta$ and $\zeta$ clearly shows that, even
for a Gaussian-distributed $\zeta$,
the corresponding matter density field is non-Gaussian.
Assuming simple inflationary Gaussian initial conditions in $\zeta$,
and using a peak-background split, we have derived the corresponding specific general-relativistic effective non-Gaussianity
parameters, \eref{eff:fNL} and \eref{gen:fNL}, that results when a Newtonian treatment is used, i.e. a Poisson equation as relation between $\delta$ and the Newtonian potential $\phi$, and a linear relation is assumed between $\zeta$ and $\phi$.
Although Newtonian simulations are commonly used to study the
non-linear evolution of the matter density contrast $\delta$, a
relativistic approach is essential to properly determine the initial
conditions set
by a period of inflation in the very early
universe. Thus we have shown how Einstein's gravity imprints a characteristic signature
in the large-scale structure of our universe.
\newpage
\section{Acknowledgements}
\label{sec:thanks}
\noindent The authors are grateful to Rob Crittenden, Roy Maartens and Gianmassimo Tasinato for useful
{discussions}.
This work was supported by STFC grants ST/K00090X/1 and ST/L005573/1, and by PAPIIT-UNAM grants IN103413-3 and IA101414-1.
|
\section{Introduction}
Quantum spin-1/2 two-leg ladders have attracted much attention from both experimental and theoretical standpoints because of their rich physical properties. At zero magnetic field, inelastic neutron-scattering experiments on these materials\cite{Hong06:74,Notbohm07:98,Savici09:80,Schmidiger11:84,Schmidiger13:88} have provided comprehensive knowledge on one-magnon and unconventional two-magnon excitations\cite{Barnes47:93,Bouillot11:83,Normand11:83} predicted by theory. In magnetic fields,\cite{Chitra55:97,Giamarchi59:99} they exhibit novel quantum-critical behavior \cite{Sachdev99} such as Bose-Einstein condensation,\cite{Garlea07:98,Thielemann09:79} magnetic Bose glass,\cite{Hong10:81} and Tomonaga-Luttinger liquid phases.\cite{Thiel09:102,Hong10:105,Ninios12:108,Schmidiger12:108,Schmidiger13:111} In addition, a quantum phase transition is expected to occur from a quantum-disordered state to a magnetically-ordered state, as the strength of interladder couplings is increased.\cite{Troyer97:55,Normand97:56,Capriotii02:65}
To date, there have been very few detailed experimental studies of coupled spin ladders because of a scarcity of suitable model systems. Na$_2$Co$_2$(C$_2$O$_4$)$_3$(H$_2$O)$_2$ (Ref.~\onlinecite{Honda05:95}) and CaCu$_2$O$_3$ (Ref.~\onlinecite{Wolf05}) were previously identified as such systems on the basis of magnetization measurements. However, further investigation with neutron scattering\cite{Matsuda07:75} has found that Na$_2$Co$_2$(C$_2$O$_4$)$_3$(H$_2$O)$_2$ is a system of almost isolated dimers, with a singlet ground state, not of a ladder. In the case of CaCu$_2$O$_3$, neutron-diffraction work\cite{Kiryukhin01:63} has revealed an incommensurate spiral magnetic structure, which originates from frustrated interladder couplings, making the system more complicated than originally thought. Interladder couplings have also been reported in the two-leg spin-ladder compounds IPA-CuCl$_3$ (Refs.~\onlinecite{Masuda06:96,Fischer11:96}) and BiCu$_2$PO$_6$ (Ref.~\onlinecite{Plumb13:88}), but their ground states remain spin liquids owing to frustrated terms in their Hamiltonian. Three-dimensional (3D) long-range ordering (LRO) has been observed at $\sim$120 K in another ladder compound, LaCuO$_{2.5}$ (Ref.~\onlinecite{Kadono96:54}), but the details of the magnetic structure and the spin dynamics of this material are still unknown.
In the present paper, we report specific-heat measurements, and neutron-diffraction and inelastic neutron-scattering (INS) experiments on what we shall show to be a coupled ladder case, (dimethylammonium)(3,5-dimethylpyridinium)CuBr$_4$, which we call DLCB. This study reveals that a spin gap coexists in this material with a 3D LRO. We also determine main exchange interactions and interaction anisotropy from the measured dispersions both along and perpendicular to the ladder directions. The ratio $\alpha$=$J_{\rm int}$/$J_{\rm leg}$ of the interladder exchange, $J_{\rm int}$, to the leg exchange, $J_{\rm leg}$, indicates that DLCB is very close to a quantum critical point at which the LRO vanishes.
The crystal structure of DLCB is triclinic, space group $P\bar{1}$, and the lattice constants at 85 K are $a=7.459$ \AA, $b=8.270$ \AA, $c=13.720$ \AA, $\alpha=107.41^\circ$, $\beta=90.21^\circ$, and $\gamma=91.37^\circ$.\cite{Awwadi08:47} Nearest-neighbor and next-nearest-neighbor contacts between bromine atoms suggest that CuBr$_4^{-2}$ anions form two-leg ladders along the crystallographic $\bm{b}$ axis as shown in Fig.~\ref{fig1}(a).~$J_{\rm leg}$\,=\,0.685\,meV and a value of the rung exchange, $J_{\rm rung}$\,=\,0.351\,meV, have been obtained from magnetic susceptibility, which shows no evidence for LRO down to 2 K (Ref.~\onlinecite{Awwadi08:47}). But different from (Hpip)$_2$CuBr$_4$ (BPCB)\cite{Savici09:80,Thielemann09:79} and (2,3-dmpyH)$_2$CuBr$_4$ (DIMPY)\cite{Hong10:105,Schmidiger12:108,Shapiro07:952}, in which ladders are well separated from each other and thus interladder couplings ($\sim$$\mu$eV) are negligible, the interladder Cu-Cu distances in DLCB are comparable to or even shorter than intraladder Cu-Cu distances. Based on the crystal structure, we propose a minimal coupling between ladders as depicted in Fig.~\ref{fig1}(b), which shows the magnetic interactions between Cu$^{2+}$ ions in the crystallographic \emph{ac} plane. The red bond is $J_{\rm rung}$ and the yellow one is the nearest-neighbor $J_{\rm int}$ along the [10$\bar{1}$] direction in real space. As we shall show below, this two-dimensional (2D) model of coupled spin ladders fully accounts for experimentally observed magnon dispersions.
\begin{figure}[htbp]
\includegraphics[width=7.0cm,bbllx=50,bblly=65,bburx=555,bbury=675,angle=-90,clip=]{fig1.ps}
\caption{(Color online) (a) Crystal structure of DLCB, in which the ladder chain extends along the $\bm{b}$ axis. Outlined is a nuclear unit cell. (b) Projection of CuBr$_4^{-2}$ tetrahedra onto a plane perpendicular to the $\bm{b}$ axis, showing proposed interladder couplings. Different lines stand for the different bonds. Red and yellow lines indicate the intraladder coupling $J_{\rm rung}$ and interladder coupling $J_{\rm int}$, respectively. Black arrows indicate the directions of the spins. Parallelogram is a projection of a magnetic unit cell.} \label{fig1}
\end{figure}
\section{Experimental Methods}
Single crystals of deuterated DLCB were grown according to the procedure described in Ref.~\onlinecite{Awwadi08:47}. The crystal structure was determined at 4 K on the four-cycle diffractometer (HB3A) at the High Flux Isotope Reactor (HFIR), Oak Ridge National Laboratory.
The magnetic neutron-diffraction measurements were made on a 0.2 g single crystal with a 0.4$^\circ$ mosaic spread, on a thermal neutron triple-axis spectrometer (HB1A) at the HFIR, with a neutron energy of 14.7 meV. A pyrolytic graphite (PG) (002) monochromator and analyzer were used together with horizontal collimation of 48$^\prime$-48$^\prime$-40$^\prime$-80$^\prime$. Contamination from higher-order beams was removed using PG filters. The sample was oriented horizontally in the (HK$\bar{\rm K}$) reciprocal-lattice plane and a continuous-flow helium-3 cryostat was used.
Inelastic neutron-scattering measurements were performed on a cold neutron triple-axis spectrometer (CTAX) at the HFIR and on a multi-angle crystal spectrometer (MACS)\cite{Rodriguez08:19} and a disk chopper time-of-flight spectrometer (DCS)\cite{Copley03} at the NIST Center for Neutron Research, on two co-aligned single crystals with a total mass of 1.0 g and a 1.0$^\circ$ mosaic spread. The sample was aligned in the either (HK$\bar{\rm H}$) or (HK0) scattering plane and standard helium-4 cryostats were used. At CTAX, the final neutron energy was set to either 3.5 or 5.0 meV by a PG (002) analyzer. The horizontal collimation was guide-open-80$^\prime$-open. At MACS, the final neutron energy was set to either 3.0 or 5.0 meV. Higher-order reflections were removed by a cooled beryllium filter placed between the sample and the analyzer. At DCS, a disk chopper was used to select a 167-Hz pulsed neutron beam with 3.27 meV. Reduction and analysis of the data were performed by using the software DAVE.\cite{Azuah09}
The specific heat was measured above 0.5 K by utilizing a commercial setup (Quantum Design, Physical Property Measurement System)\cite{ppms} and below 0.5 K in a home-made calorimeter\cite{Tsujii} in a dilution refrigerator.
\section{Experimental Results}
To investigate the ground state of DLCB, we first performed specific-heat measurements, whose result is shown in Fig.~\ref{fig2}(a). The sharp anomaly at about 2.0 K indicates a phase transition to a LRO state. In addition, the exponential, activated behavior at low temperatures, shown in the inset to the figure, reveals the presence of a spin gap. By fitting the data to a formula for the specific heat of a one-dimensional $S$=1/2 gapped Heisenberg antiferromagnet in the low-temperature limit,\cite{Troyer94:50}
\begin{eqnarray}
C_{m} (T)\propto\left(\frac{\Delta}{k_B T}\right)^{3/2}\Delta e^{-\Delta/k_B T}, \label{eq:ct1}
\end{eqnarray}
we find the energy gap $\Delta$=0.29(2) meV. In a conventional $S$=1/2 two-leg spin-ladder system, the presence of a spin gap is tantamount to the ground state being quantum disordered. It might therefore seem surprising that a spin gap coexists with LRO in DLCB. We return later to this counterintuitive result.
\begin{figure}[htbp]
\includegraphics[width=7.0cm,bbllx=0,bblly=15,bburx=550,bbury=720,angle=0,clip=]{fig2.ps}
\caption{(Color online) (a) Specific heat of DLCB. Inset: semilog plot of the magnetic component of specific heat, after subtraction of a phonon contribution, against $1/T$. The solid line is a fit showing exponential, activated behavior at low temperatures. (b) Background-subtracted neutron-peak intensity at $\bm{q}$=(0.5,0.5,$-$0.5) as a function of temperature. Error bars represent one standard deviation. The solid line is a fit to a power law as described in the text. Inset: rocking-curve scans through $\bm{q}$=(0.5,0.5,$-$0.5), made at \emph{T}=0.3 and 2.5 K. The solid line is a guide to the eye.} \label{fig2}
\end{figure}
To further examine the ordered ground state, we carried out single-crystal neutron diffraction measurements. At 0.3 K, we have collected in total 26 nuclear and 32 magnetic reflections, from which the magnetic propagation vector was found to be (0.5,0.5,0.5). The data were analyzed by the Rietveld method using the FULLPROF program.\cite{fullprof} The resulting collinear spin structure is depicted in Fig.~\ref{fig1}(b), with alternating moments pointing along the $\bm{c}^*$ axis in the reciprocal-lattice space. Not shown in the figure, the directions of the moments also alternate along the ladder. The size of the local moment is only 0.39(5)\,$\mu_B$ even at 0.3 K, much smaller than $\mu_B$, due to quantum fluctuations.
The inset to Fig.~\ref{fig2}(b) shows representative $\theta$ scans through $\bm{q}$=(0.5,0.5,$-$0.5) measured at 0.3 and 2.5 K. The scan at 0.3 K shows a peak, which disappears as the temperature is raised above about 2.0 K, thus confirming its magnetic origin. The temperature dependence of the intensity of the Bragg peak at $\bm{q}$=(0.5,0.5,$-$0.5) is plotted in Fig.~\ref{fig2}(b). A power-law fit of the form ($1-T/T_N$)$^{2\beta}$ gives a $T_N$ of 1.99(2)\,K, in good agreement with the specific-heat data, and the critical exponent $\beta$=0.28(2), which is smaller than $\beta$$\simeq$0.37 and 0.33 for the 3D Heisenberg and Ising universality classes, respectively.\cite{Peli}
INS was used to study the spin dynamics in DLCB. For all the results presented here, data taken at 10\,K with the same instrument configuration has been subtracted as a background. Figure~\ref{fig3}(a) shows a constant-$\bm{q}$ scan at the magnetic zone center (0.5,0.5,$-$0.5) at \emph{T}=1.5\,K. The instrumental-resolution-limited peak indicates a spin gap of 0.30(2) meV, in excellent agreement with the specific-heat result.
Figures~\ref{fig3}(b)--\ref{fig3}(e) show the evolution of the spin wave at 1.7 K in a few Brillouin zones (BZ) with increasing energy $\hbar\omega$. The magnetic intensity develops above the spin gap. As indicated by the ellipse-like shape of the intensity profile in Figs.~\ref{fig3}(b) and \ref{fig3}(c), the dispersion is weaker along [10$\bar{1}$], perpendicular to the ladder direction, than along the ladder direction [010]. The intensity disappears at $\hbar\omega$$\sim$0.8 meV for the dispersion perpendicular to the ladder and at about 1.3 meV for the dispersion along the ladder, as shown in Figs.~\ref{fig3}(d) and \ref{fig3}(e).
\begin{figure}[htbp]
\includegraphics[width=8cm,bbllx=35,bblly=85,bburx=575,bbury=765,angle=0,clip=]{fig3.ps}
\caption{(Color online) (a) Background-subtracted constant-\textbf{\emph{q}} scan in DLCB at the magnetic zone center (0.5,0.5,{$-$}0.5), measured at CTAX at \emph{T}=1.5 K. The solid line is a fit to a Gaussian profile convolved with the instrumental resolution function. Error bars represent one standard deviation. (b--e) Background-subtracted constant-energy slices measured at MACS at $T$=1.7 K for excitation energies of 0.2--0.4 meV, 0.4--0.6 meV, 0.7--0.9 meV, and 1.2--1.4 meV.} \label{fig3}
\end{figure}
The spin-wave dispersions along these high-symmetry directions, [010] and [10$\bar{1}$], are presented in Figs.~\ref{fig4}(a) and \ref{fig4}(b), respectively. The bandwidths of the acoustic branch for these directions are 0.82(3) and 0.35(3) meV. As a result of the LRO, the magnetic unit cell is twice as large as the nuclear unit cell along the ladder direction. Consequently, the BZ is reduced to one half, and 0.25 and 0.75 become the zone boundaries. The observed spectrum termination due to this BZ folding is similar to that in the field-induced ordered phase of IPA-CuCl$_3$ (Ref.~\onlinecite{Zheludev07:76}). Furthermore, the presence of a flat optical branch as shown in Fig.~\ref{fig4}(c) arises from the alternation of $J$ along the [10$\bar{1}$] direction.
\begin{figure}[htbp]
\includegraphics[width=7.0cm,bbllx=20,bblly=30,bburx=500,bbury=790,angle=0,clip=]{fig4.ps}
\caption{(Color online) False-color map of the spin-wave spectra at $T$=1.5 K (a) along the ladders [010] and (b) perpendicular to the ladders along [10$\bar{1}$] measured at DCS, and (c) perpendicular to the ladders along [100] measured at MACS. Red and yellow lines are calculations for the acoustic and optical branches of one-magnon dispersion, respectively. Data shown as circles were obtained at CTAX.} \label{fig4}
\end{figure}
\section{Analysis and Discussion}
One peculiarity of the LRO in DLCB is that it does not produce a linear, gapless Nambu-Goldstone mode. This absence must arise from an inherent easy-axis anisotropy, which reduces the symmetry of the system to an axial one. In the ordered state, the spins form a collinear structure as shown in Fig.~\ref{fig1}(b), a structure which does not break the axial symmetry, hence the absence of a Nambu-Goldstone mode. This reasoning is borne out by detailed analysis of the spin-wave dispersions, as described below.
The simplest Hamiltonian that accounts for this scenario is
\begin{eqnarray} \label{ladeqn}
H &=& \sum_{\gamma, \langle i,j\rangle}J_{\gamma} \left[ S^{\rm z}_{i}S^{\rm z}_{j}+\lambda\left( S^{\rm x}_{i} S^{\rm x}_{j}+S^{\rm y}_{i}S^{\rm y}_{j}\right)\right],
\end{eqnarray}
where the subscript $\gamma$ reads either `rung', `leg', or `int'---for $J_{\gamma}$ being the rung, leg, or interladder exchange constant---and $i$ and $j$ are nearest-neighbor lattice sites. The parameter $\lambda$ specifies an interaction anisotropy, with $\lambda$=0 and 1 being the limiting cases of Ising and Heisenberg interactions, respectively. We assume that $\lambda$ is the same for all three $J_{\gamma}$'s in order to minimize the number of fitting parameters to be determined from our spin-wave dispersion data.
To calculate the dispersion for this 2D model, we first perform a sublattice rotation, which transforms Eq.~(\ref{ladeqn}) to
\begin{equation} \label{ham_rotated}
H = \sum_{\gamma, \langle i,j\rangle} J_{\gamma} \left[-S^{\rm z}_{i} S^{\rm
z}_{j}-\frac{\lambda}{2}\left(S^{\rm +}_{i}S^{\rm +}_{j}+S^{\rm -}_{i}S^{\rm -}_{j}\right)\right].
\end{equation}
We then introduce hardcore-boson operators $\hat{a}^{\dagger}_\nu$ and $\hat{a}^{\phantom{\dagger}}_\nu$ ($\hat{b}^{\dagger}_\nu$ and $\hat{b}^{\phantom{\dagger}}_\nu$), which create and annihilate a magnon at site $A$ ($B$) belonging to rung $\nu$ in the ferromagnetic reference state, obtaining the hardcore-boson Hamiltonian
\begin{eqnarray}\label{ham_bosons}
\frac{\hat{H}}{\tilde{J}} &=& -\frac{N}{2}+\sum_{\nu}\left( \hat{n}_{\nu}^{(a)}+\hat{n}_{\nu}^{(b)}\right)\nonumber\\
&& +\sum_\nu \left[ \hat{T}_{\nu,0}+\lambda\left( \hat{T}_{\nu,-2}+\hat{T}_{\nu,+2}\right)\right]\quad ,
\end{eqnarray}
where $\tilde{J}=J_{\rm leg}+(J_{\rm rung}+J_{\rm int})/2$, $N$ is the number of unit cells, $\hat{n}_{\nu}^{(a)}=\hat{a}^\dagger_\nu\hat{a}^{\phantom{\dagger}}_\nu$, and $\hat{n}_{\nu}^{(b)}=\hat{b}^\dagger_\nu\hat{b}^{\phantom{\dagger}}_\nu$. The sums are taken over all rungs. The operators $\hat{T}_{\nu,n}$, with $\hat{T}_{\nu,-2}=\hat{T}_{\nu,+2}^\dagger$, are given by
\begin{eqnarray}\label{T0}
\hat{T}_{\nu,0} &=& -x_{\rm rung} \hat{n}_{\nu}^{(a)}\hat{n}_{\nu}^{(b)}-x_{\rm int} \hat{n}_{\nu}^{(b)}\hat{n}_{\nu+\delta x}^{(a)}\nonumber\\
&& -x_{\rm leg} \left( \hat{n}_{\nu}^{(a)}\hat{n}_{\nu+\delta y}^{(a)} + \hat{n}_{\nu}^{(b)}\hat{n}_{\nu+\delta y}^{(b)}\right)
\end{eqnarray}
and
\begin{eqnarray}\label{T+2}
\hat{T}_{\nu,+2} &=& -x_{\rm rung} \hat{a}^\dagger_{\nu}\hat{b}^\dagger_{\nu}-x_{\rm int} \hat{b}^\dagger_{\nu}\hat{a}^\dagger_{\nu+\delta x}\nonumber\\
&& -x_{\rm leg} \left( \hat{a}^\dagger_{\nu}\hat{a}^\dagger_{\nu+\delta y} + \hat{b}^\dagger_{\nu}\hat{b}^\dagger_{\nu+\delta y}\right) \quad ,
\end{eqnarray}
where $x_{\gamma}=J_{\gamma}/\tilde{J}$, and $\delta x$ ($\delta y$) is the distance between two neighboring rungs belonging to different ladders (the same ladder).
Finally, perturbative continuous unitary transformations\cite{Knetter2000,Knetter2003} map Eq.~(\ref{ham_bosons}) to an effective model, $H_{\rm eff}$, which conserves the number of magnons. The one-magnon sector $H^{\rm (1)}_{\rm eff}$, which is of our interest, can be simplified by Fourier transform to
\begin{equation}
H^{\rm (1)}_{\rm eff} = \sum_{\bm{k}} \left( \omega_{\alpha}(\bm{k}) \alpha^\dagger_{{\bm{k}}} \alpha^{\phantom{\dagger}}_{\bm{k}}+ \omega_{\beta}(\bm{k}) \beta^\dagger_{{\bm{k}}} \beta^{\phantom{\dagger}}_{\bm{k}} \right) \quad ,
\end{equation}
where $\omega_{\alpha}(\bm{k})$ and $\omega_{\beta}(\bm{k})$ denote the two one-magnon branches stemming from the two-site unit cell. They are calculated as follows.
We first replace $x_\gamma$ in Eqs.~(\ref{T0}) and (\ref{T+2}) with $\tau x_\gamma$, so that at $\tau$=0, Eq.~(\ref{ham_bosons}) will contain only the first two terms. This is our unperturbed Hamiltonian. The perturbation expansion then gives $\omega_{\alpha}(\bm{k})$ and $\omega_{\beta}(\bm{k})$ as power series of $\tau$. We next express $\omega_{\alpha}^{2}(\bm{k})$ and $\omega_{\beta}^{2}(\bm{k})$ as Pad\'{e} approximants $P_{l}(\tau)/Q_{m}(\tau)$, where $P_{l}(\tau)$ and $Q_{m}(\tau)$ are power polynomials of order $l$ and $m$, respectively. The approximants are uniquely determined by choosing $l+m=n$, where $n$ is the order of the raw perturbation series.
The reason for casting $\omega_{\alpha}^{2}(\bm{k})$ and $\omega_{\beta}^{2}(\bm{k}),$ instead of $\omega_{\alpha}(\bm{k})$ and $\omega_{\beta}(\bm{k})$, as Pad\'{e} approximants is that the one-magnon energy gap \mbox{$\Delta=\omega(0,0)$} of a square-lattice antiferromagnet vanishes at $\lambda=1$ with a square-root singularity.\cite{Zheng91} In other words~$\Delta^2$ is a linear function~of $\lambda$ close to the critical point. This behavior of $\Delta^2$ is ensured by expressing $\omega_{\alpha}^{2}(\bm{k})$ as a Pad\'{e} approximant.
Finally, we set $\tau$=1 to obtain $\omega_{\alpha}^2(\bm{k})$ and $\omega_{\beta}^2(\bm{k})$ for the full Hamiltonian, Eq.~(\ref{ham_bosons}). We have found that perturbation expansion up to order $n=13$ results in
theoretical uncertainties that are well below the experimental error bars for all the data points shown in Fig.~\ref{fig4}.
To compare the calculation with the experimental data shown in Fig.~\ref{fig4}, we (i) shift the two spin-wave dispersions $\omega_{\alpha}(\bm{k})$ and $\omega_{\beta}(\bm{k})$ by $\bm{k}=(0,\pi)$ and (ii) fold the resulting dispersions into the reduced BZ in the $y$ direction, because the unit cell of the N\'eel-ordered state comprises four sites before the sublattice rotation. We choose the exchange constants $J_\gamma$ and $\lambda$ such that the calculated curves lie within experimental error bars. The results are in excellent agreement with the observed dispersions, as shown in Fig.~\ref{fig4}, yielding $J_{\rm leg}=0.60(2)$\,meV, $J_{\rm rung}=0.64(9)$\,meV, $J_{\rm int}=0.19(2)$\,meV, and $\lambda=0.93(2)$. The value of $J_{\rm leg}$ is somewhat smaller than 0.685 meV determined from magnetic susceptibility, whereas $J_{\rm rung}$ is much larger than 0.351 meV from the susceptibility.\cite{Awwadi08:47} The discrepancy is not surprising, given that the susceptibility analysis has assumed that $J_{\rm int}$ is negligible.
The exchange ratios $x$=$J_{\rm leg}/J_{\rm rung}$ and $\alpha$=$J_{\rm int}/J_{\rm leg}$ are~0.94(14) and~0.32(3), respectively. For Heisenberg $S$=1/2 coupled two-leg square ($x$=1) ladders, the quantum critical point between the spin-liquid phase and the N\'eel-ordered phase has been predicted to be at $\alpha_c$$\sim$0.3.\cite{Capriotii02:65} Since difference of the energy gaps of isolated spin ladders with $x$=0.94 and $x$=1 is as small as 0.01$J_{\rm rung}$, which is negligible,\cite{Hong10:105} we expect $\alpha_c$ for $x$=0.94 to be $\sim$0.3/0.94=0.32, which turns out to be the same as $\alpha$ for DLCB within our uncertainties. Thus DLCB is an ideal experimental realization of a coupled spin-ladder system in which the ground state becomes magnetically ordered with an $\alpha$ very close to the critical value. The weak Ising-like anisotropy ($\lambda$=0.93) prevents the gap from closing, while allowing---and to some extent even promoting---the ordering at finite temperature even if the interlayer coupling is absent.
\section{Conclusion}
In summary, we have carried out specific-heat and neutron-scattering measurements on DLCB to determine its spin Hamiltonian and the ground state. We have found a long-range magnetic order coexisting with a spin energy gap. An easy-axis anisotropy, which accounts for the coexistence, has also been identified. The measured magnetic dispersions are quantitatively consistent with a coupled $S$=1/2 two-leg spin-ladder model with ladder legs along the $\bm{b}$ direction.
\begin{acknowledgments}
TH thanks D. A. Tennant for a helpful discussion. We thank J.-H. Park and G. E. Jones for help with cryogenics. The work at the HFIR, Oak Ridge National Laboratory, was sponsored by the Division of Scientific User Facilities, Office of Basic Energy Science, US Department of Energy (DOE). The work at NIST utilized facilities supported by the NSF under Agreement Nos. DMR-9986442, -0086210, and -0454672. The National High Magnetic Field Laboratory (NHMFL), in which the specific-heat measurements were made, is supported by NSF Cooperative Agreement No. DMR-0654118, by the State of Florida, and by the DOE. XK acknowledges support from Michigan State University, CPA and YT acknowledge support by the NHMFL UCGP program, and HDZ acknowledges support from the JDRD program of the University of Tennessee.
\end{acknowledgments}
\thebibliography{}
\bibitem{Hong06:74} T. Hong, M. Kenzelmann, M. M. Turnbull, C. P. Landee, B. D. Lewis, K. P. Schmidt, G. S. Uhrig, Y. Qiu, C. Broholm, and D. Reich, Phys. Rev. B {\bf{74}}, 094434 (2006).
\bibitem{Notbohm07:98} S. Notbohm, P. Ribeiro, B. Lake, D. A. Tennant, K. P. Schmidt, G. S. Uhrig, C. Hess, R. Klingeler, G. Behr, B. B\"{u}chner, M. Reehuis, R. I. Bewley, C. D. Frost, P. Manuel, and R. S. Eccleston, Phys. Rev. Lett. {\bf{98}}, 027403 (2007).
\bibitem{Savici09:80} A. T. Savici, G. E. Granroth, C. L. Broholm, D. M. Pajerowski, C. M. Brown, D. R. Talham, M. W. Meisel, K. P. Schmidt, G. S. Uhrig, and S. E. Nagler, Phys. Rev. B {\bf{80}}, 094411 (2009).
\bibitem{Schmidiger11:84} D. Schmidiger, S. M$\rm\ddot{u}$hlbauer, S. N. Gvasaliya, T. Yankova, and A. Zheludev, Phys. Rev. B {\bf{84}}, 144421 (2011).
\bibitem{Schmidiger13:88} D. Schmidiger, S. M$\rm\ddot{u}$hlbauer, A. Zheludev, P. Bouillot, T.~Giamarchi, C.~Kollath, G. Ehlers, and A. M. Tsvelik, Phys. Rev. B {\bf{88}}, 094411 (2013).
\bibitem{Barnes47:93} T. Barnes, E. Dagotto, J. Riera, and E. S. Swanson, Phys. Rev. B {\bf{47}}, 3196 (1993).
\bibitem{Bouillot11:83} P. Bouillot, C. Kollath, A. M. L\"{a}uchli, M. Zvonarev, B. Thielemann, Ch. R$\rm\ddot{u}$egg, E. Orignac, R. Citro, M. Klanj\v{s}ek, C. Berthier, M. Horvati$\rm\acute{c}$, and T. Giamarchi, Phys. Rev. B {\bf{83}}, 054407 (2011).
\bibitem{Normand11:83} B. Normand and Ch. R$\rm\ddot{u}$egg, Phys. Rev. B {\bf{83}}, 054415 (2011).
\bibitem{Chitra55:97} R. Chitra and T. Giamarchi, Phys. Rev. B. {\bf{55}}, 5816 (1997).
\bibitem{Giamarchi59:99} T. Giamarchi and A. M. Tsvelik, Phys. Rev. B. {\bf{59}}, 11398 (1999).
\bibitem{Sachdev99} S. Sachdev, \emph{Quantum Phase Transition} (Cambridge University Press\, Cambridge, 1999).
\bibitem{Garlea07:98} V.~O.~Garlea, A.~Zheludev, T.~Masuda, H.~Manaka, L.-P.~Regnault, E.~Ressouche, B.~Grenier, J.-H.~Chung, Y.~Qiu, K.~Habicht, K.~Kiefer, and M.~Boehm, Phys. Rev. Lett. {\bf 98}, 167202 (2007).
\bibitem{Thielemann09:79} B. Thielemann, Ch. R$\rm \ddot{u}$egg, K. Kiefer, H. M. R{\o}nnow, B. Normand, P. Bouillot, C. Kollath, E. Orignac, R. Citro, T. Giamarchi, A. M. L$\rm \ddot{a}$uchli, D. Biner, K. W. Kr\"{a}mer, F. Wolff-Fabirs, V. S. Zapf, M. Jaime, J. Stahn, N. B. Christensen, B. Grenier, D. F. McMorrow, and J. Mesot, Phys. Rev. B {\bf 79}, 020408(R) (2009).
\bibitem{Hong10:81} T.~Hong, A.~Zheludev, H.~Manaka, and L.-P.~Regnault, Phys. Rev. B {\bf{81}}, 060410 (2010).
\bibitem{Thiel09:102} B.~Thielemann, Ch.~R$\rm \ddot{u}$egg, H.~M.~R$\rm {\o}$nnow, A.~M.~L$\rm \ddot{a}$uchli, J.-S.~Caux, B.~Normand, D.~Biner, K.~W.~Kr$\rm \ddot{a}$mer, H.-U.~G$\rm \ddot{u}$del, J.~Stahn, K.~Habicht, K.~Kiefer, M.~Boehm, D.~F.~McMorrow, and J.~Mesot, Phys. Rev. Lett. {\bf 102}, 107204 (2009).
\bibitem{Hong10:105} T.~Hong, Y.~H.~Kim, C.~Hotta, Y.~Takano, G.~Tremelling, M.~M.~Turnbull, C.~P.~Landee, H.-J.~Kang, N.~B.~Christensen, K.~Lefmann, K.~P.~Schmidt, G.~S.~Uhrig, and C.~Broholm, Phys. Rev. Lett. {\bf 105}, 137207 (2010)
\bibitem{Ninios12:108} K.~Ninios, T.~Hong, T.~Manabe, C.~Hotta, S.~N.~Herringer, M.~M.~Turnbull, C.~P.~Landee, Y.~Takano, and H.~B.~Chan, Phys. Rev. Lett. {\bf 108}, 097201 (2012).
\bibitem{Schmidiger12:108} D. Schmidiger, P. Bouillot, S. M\"{u}hlbauer, S. Gvasaliya, C.~Kollath, T.~Giamarchi, and A. Zheludev, Phys. Rev. Lett. {\bf{108}}, 167201 (2012).
\bibitem{Schmidiger13:111} D. Schmidiger, P. Bouillot, T.~Guidi, R.~Bewley, C.~Kollath, T.~Giamarchi, and A. Zheludev, Phys. Rev. Lett. {\bf{111}}, 107202 (2013).
\bibitem{Normand97:56} B. Normand and T. M. Rice, Phys. Rev. B {\bf{56}}, 8760 (1997).
\bibitem{Troyer97:55} M. Troyer, M. E. Zhitomirsky, and K. Ueda, Phys. Rev. B {\bf{55}}, R6117 (1997).
\bibitem{Capriotii02:65} L. Capriotti and F. Becca, Phys. Rev. B {\bf{65}}, 092406 (2002), and references therein.
\bibitem{Honda05:95} Z. Honda, K. Katsumata, A. Kikkawa, and K. Yamada, Phys. Rev. Lett. {\bf{95}}, 087204 (2005).
\bibitem{Wolf05} M. Wolf, K.-H. M$\rm\ddot{u}$ller, D. Eckert, S.-L. Drechsler, H. Rosner, C. Sekar, and G. Krabbes, J. Magn. Magn. Mater. {\bf{290-291}}, 314 (2005).
\bibitem{Matsuda07:75} M. Matsuda, S. Wakimoto, K. Kakurai, Z. Honda, and K. Yamada, Phys. Rev. B {\bf{75}}, 012405 (2007).
\bibitem{Kiryukhin01:63} V. Kiryukhin, Y. J. Kim, K. J. Thomas, F. C. Chou, R. W. Erwin, Q. Huang, M. A. Kastner, and R. J. Birgeneau, Phys. Rev. B {\bf{63}}, 144418 (2001).
\bibitem{Masuda06:96} T. Masuda, A. Zheludev, H. Manaka, L.-P. Regnault, J.-H. Chung, and Y. Qiu, Phys. Rev. Lett. {\bf{96}}, 047210 (2006).
\bibitem{Fischer11:96} T. Fischer, S. Duffe, and G. S. Uhrig, EPL {\bf{96}}, 47001 (2011).
\bibitem{Plumb13:88} K. W. Plumb, Z. Yamani, M. Matsuda, G. J. Shu, B. Koteswararao, F. C. Chou, and Y.-J. Kim, Phys. Rev. B {\bf{88}}, 024402 (2013).
\bibitem{Kadono96:54} R. Kadono, H. Okajima, A. Yamashita, K. Ishii, T. Yokoo, J. Akimitsu, N. Kobayashi, Z. Hiroi, M. Takano, and K. Nagamine, Phys. Rev. B {\bf{54}}, R9628(R) (1996).
\bibitem{Awwadi08:47} F. Awwadi, R. D. Willett, B. Twamley, R. Schneider, and C. P. Landee, Inorg. Chem. {\bf{47}}, 9327 (2008).
\bibitem{Shapiro07:952} A. Shapiro, C. P. Landee, M. M. Turnbull, J. Jornet, M. Deumal, J. J. Novoa, M. A. Robb, and W. Lewis, J. Am. Chem. Soc. {\bf 129}, 952 (2007).
\bibitem{Rodriguez08:19} J. A. Rodriguez, D. M. Adler, P. C. Brand, C. Broholm, J. C. Cook, C. Brocker, R. Hammond, Z. Huang, P. Hundertmark, J. W. Lynn, N. C. Maliszewskyj, J. Moyer, J. Orndorff, D. Pierce, T. D. Pike, G. Scharfstein, S. A. Smee, and R. Vilaseca, Meas. Sci. Technol. {\bf{19}}, 034023 (2008).
\bibitem{Copley03} J. R. D. Copley and J. C. Cook, Chem. Phys. {\bf{292}}, 477 (2003).
\bibitem{Azuah09} R. T. Azuah, L. R. Kneller, Y. Qiu, P. L. W. Tregenna-Piggott, C. M. Brown, J. R. D. Copley, and R. M. Dimeo, J. Res. Natl. Inst. Stand. Technol. {\bf{114}}, 341 (2009).
\bibitem{ppms} The identification of certain commercial products and their suppliers should in no way be construed as indicating that such products or suppliers are endorsed by NIST or are recommended by NIST or that they are necessarily the best for the purposes described.
\bibitem{Tsujii} H. Tsujii, B. Andraka, E. C. Palm, T. P. Murphy, and Y. Takano, Physica\ B \textbf{329--333}, 1638 (2003).
\bibitem{Troyer94:50} M.~Troyer, H.~Tsunetsugu, and D.~W$\rm\ddot{u}$rtz, Phys. Rev. B {\bf 50}, 13515 (1994).
\bibitem{fullprof} J. Rodriguez-Carvajal, Physica B {\bf 192}, 55 (1993).
\bibitem{Peli} A. Pelissetto and E. Vicari, Phys. Rep. {\bf{368}}, 549 (2002), and references therein.
\bibitem{Zheludev07:76} A.~Zheludev, V.~O.~Garlea, T.~Masuda, H.~Manaka, L.-P.~Regnault, E.~Ressouche, B.~Grenier, J.-H.~Chung, Y.~Qiu, K.~Habicht, K.~Kiefer, and M.~Boehm, Phys. Rev. B {\bf 76}, 054450 (2007).
\bibitem{Knetter2000}
C. Knetter and G. S. Uhrig, Eur. Phys. J. B {\bf 13}, 209 (2000).
\bibitem{Knetter2003}
C. Knetter, K. P. Schmidt, and G. S. Uhrig, J. Phys. A: Math. Gen. {\bf 36}, 7889 (2003).
\bibitem{Zheng91} W. Zheng. J. Oitmaa, and C. J. Hamer, Phys. Rev. B {\bf 43}, 8321 (1991).
\end{document}
|
\section{Introduction}
Nearly a century has passed since Albert Einstein wrote down the field
equations of general relativity. A crucial prediction of his theory is
the existence of GWs. Observations of the Hulse-Taylor binary pulsar
\citep{Taylor:1989} and the double pulsar J0737-3039 \citep{Lyne:2004}
leave little doubt of the existence of GWs, with further evidence
provided by the recent claim of a detection of a GW-induced B-mode
polarization of the cosmic microwave
background~\citep{2014arXiv1403.3985B}. However, GWs still elude
direct observation. The situation should change in the next few
years, when a network of second-generation GW
observatories -- including Advanced LIGO (Harry, \citeyear{AdvLIGO},
henceforth aLIGO),
Advanced Virgo \citep[][henceforth AdV]{AdvVirgo}, and KAGRA
\citep{KAGRA} -- will start taking data. The unprecedented
sensitivity of these observatories will allow them to observe the
inspiral and merger of DCOs out to cosmological distances: for
example, aLIGO should observe binary neutron stars out to a luminosity
distance of $\simeq 450 \unit{Mpc}$ ($z \sim 0.1$), while DCOs
containing BHs will be observable to much larger distances
\citep[e.g.,][]{2010CQGra..27q3001A}. Given the cosmological reach of
second-generation GW interferometers, a theoretical investigation of
the observable DCO populations which incorporates cosmological
evolution and accurate models of the gravitational waveforms is
particularly timely. This is the goal of this paper, the third in a
series \cite[cf.][]{dominik,dominik2}. Our work builds on the results
presented in the second paper \citep[henceforth Paper 2]{dominik2},
where we presented the cosmological distribution of DCOs for a set of
four evolutionary models. These models investigated a range of
Hertzsprung gap (HG) common envelope (CE) donors, supernova (SN)
explosion engines, and BH natal kicks, showing distinct differences in
the properties of the resulting DCO populations. Population models
were placed in a cosmological context by adopting the star formation
history reported in \cite{strolger} and the galaxy mass distribution
of \cite{fontana}, both of which are redshift-dependent. We performed
all calculations assuming two scenarios for metallicity evolution,
meant to bracket the uncertainties associated with the chemical
composition of the Universe. Binary evolution was performed using the
{\tt StarTrack} population synthesis code \citep{startrack}.
In this work we complete and extend the analysis of Paper 2. We study
the detection rates and the expected physical properties of coalescing
DCOs at cosmological distances for second-generation GW observatories.
The rates are calculated for different sets of gravitational waveform
models and different detector sensitivities, representative of aLIGO,
AdV, and KAGRA. Several different groups have presented similar
estimates and studies in the past decade
\citep[e.g.,][]{lipunov1997,bethe,dedonder,bloom,grishchuk:2001,nele2001,voss,dewi,nutzman,pfahl,
danny,PostnovYungelson:2006,seba,mennekens}. However,
none have combined cosmological DCO populations with accurate GW
models to obtain thorough, detector-specific results.
Our astrophysical models for DCO formation are reviewed in Section
\ref{binevol}. Gravitational waveform models and signal-to-noise ratio
estimates are discussed in Section \ref{wmodels}. Our procedure to
compute event rates is presented in Section \ref{sec:fullrates}. Event
rates and bulk properties of the detected populations are presented in
Section \ref{sec:results}. In Section \ref{sec:nobhbh} we present and discuss the study
by \cite{mennekens}, the primary result of which is the lack of detectable BH-BH systems.
In Section \ref{sec:conclusions} we discuss the possible astrophysical payoff of the
first GW detections and important directions for future work.
\section{Astrophysical models} \label{binevol}
\subsection{Binary evolution}
We begin with a summary of the four {\tt StarTrack}
evolutionary models that form the backbone of this work; a more
detailed discussion can be found in \cite{dominik,dominik2}.
\smallskip
\noindent
\textit{1) Standard model}. This is our reference model, representing
the state of the art in the formation and evolution of binary
systems. We consider only field populations here. Rate estimates
performed for dense populations in which dynamical interactions
between stars are important (i.e., globular clusters and galactic nuclear clusters) have been
presented elsewhere
\citep{gultekin,oleary,grin2006,sadowski,ivan,downing,MillerLauburg:2008}. Our
Standard model uses the ``Nanjing'' \citep{chlambda} $\lambda$
coefficient in the CE energy balance prescription of \cite{webbink},
where the precise value of $\lambda$ depends on the evolutionary stage
of the donor, its Zero Age Main Sequence (ZAMS) mass, the mass of its
envelope, and its radius. In turn, these quantities depend on
metallicity, which in our simulations varies within the broad range
$10^{-4}\leq Z \leq 0.03$ (recall that solar metallicity corresponds
to ${\rm Z}_{\odot}=0.02$). The values of $\lambda$ for high-mass stars
($M_{ZAMS}>20{\rm ~M}_{\odot}$) were obtained through private communication with
the authors and are not present in \cite{chlambda}.
The impact of the CE outcome on binary populations depends strongly on the
evolutionary stage of the donor, as first discussed in \cite{rarity}. The
Standard model does not allow for CE events with HG donors. These stars are
not expected to possess a clear core-envelope structure \citep{ivanovataam},
thus making it difficult for them to eject their outer layers during the CE phase.
In our Standard model all CE events with HG donors lead to a prompt
merger before a DCO binary is formed, regardless of the
aforementioned energy balance.
The model employs a Maxwellian distribution of natal kicks for NSs
with 1-D root mean square velocity $\sigma=265$ km/s, consistent with
NS observations \citep{hobbs}. The same distribution is extended to
BHs, where we allow for the possibility that the kicks may be reduced due
to fallback of material during the SN that leads to BH formation. The
reduction in BH kicks is described via
\begin{equation} \label{vkick}
V_{\rm k}=V_{\rm max}(1-f_{\rm fb}),
\end{equation}
where $V_{\rm k}$ is the final magnitude of the natal kick, $V_{\rm max}$
is the velocity drawn from a Maxwellian kick distribution,
and $f_{\rm fb}$ is a ``fallback factor'' that depends on the amount
of fallback material, calculated according to the prescription given in
\cite{chrisija}.
Our Standard model uses the ``Rapid'' convection-driven, neutrino-enhanced SN engine \citep{chrisija}.
The SN explosion is sourced from the Rayleigh-Taylor instability and occurs within the first
$0.1\,$--$\,0.2\,\mbox{s}$ after the bounce. When used in the context of binary evolution
models, this SN engine successfully reproduces the mass gap
\citep{massgap} observed in Galactic X-ray binaries \citep{mg1,mg2},
but see also \cite{2012ApJ...757...36K}.
\smallskip
\noindent
\textit{2) Optimistic Common Envelope}. In this model we allow HG
stars to be CE donors. When the donor initiates the CE phase, the CE
outcome is determined via energy balance. The remaining physics is
identical to the Standard model.
\smallskip
\noindent
\textit{3) Delayed SN}. This model utilizes the ``Delayed'' SN engine
instead of the Rapid one. The former is also a convection driven,
neutrino enhanced engine, but is sourced from the standing accretion
shock instability (SASI), and can produce an explosion as late as
$1\,\mbox{s}$ after bounce. The Delayed engine produces a continuous
mass spectrum of compact objects, ranging from NSs through light BHs to
massive BHs \citep{massgap}.
\smallskip
\noindent
\textit{4) High BH kicks}. In this model the BHs receive full natal
kicks, i.e. we set $f_{\rm fb}=0$ in Eq.~(\ref{vkick}). Otherwise this model is
identical to the Standard model.
\subsection{Metallicity evolution}\label{sec:metallicity}
In this paper we employ two distinct metallicity evolution scenarios:
``high-end'' and ``low-end''. These are
identical to those in our previous study (Paper 2), and a detailed
description can be found therein.
Employing such calibrations allows us to explore and bracket
uncertainties in the chemical evolution of the Universe.
In both cases the average metallicity decreases with increasing
redshift.
The high-end metallicity profile is calibrated to yield a median value
of metallicity equal to $1.5\,{\rm Z}_{\odot}$ (or $8.9$ in the ``12+log(O/H)''
formalism) at redshift $z=0$. This calibration was designed to match
the upper $1 \sigma$ scatter of metallicities according to \cite{yuan}
(see their Fig.~2, top-right panel).
The low-end metallicity profile is based on SDSS observations \citep{panter},
from which we infer that one half of the star forming mass of galaxies
at $z\sim0$ has $20\%$ solar metallicity, while the other half has
$150\%$ solar metallicity.
\section{Waveform models} \label{wmodels}
\subsection{Order-of-magnitude estimates} \label{sec:simplerates}
For any given GW detector the ``horizon distance'', $D_h$,
is defined as the luminosity distance at which an optimally oriented (face-on, overhead)
canonical $(1.4+1.4)~M_\odot$ NS-NS binary would be detected at a
fiducial threshold signal-to-noise ratio (SNR), taken to be $8$ in this paper. The
expectation value of the SNR, $\rho$, of a signal with GW amplitude $h(t)$ is given by
\begin{equation}\label{SNR}
\rho^2 = 4\int_0^\infty \frac{|\tilde h(f)|^2}{S_n(f)} df\,,
\end{equation}
where $\tilde h(f)$ is the Fourier transform of the signal and
$S_n(f)$ is the noise power spectral density of the detector
\citep[see e.g.][]{cutlerflanagan,poissonwill}. The square root of the
noise power spectral density is plotted in Fig.~\ref{fig:noise} for
several advanced interferometers of interest. For example, the aLIGO
horizon distance is $D_{h} \simeq 450 \unit{Mpc}$.
Although the sensitivity of a GW detector network depends
on the details of the search pipeline and the detector data quality, we
follow \cite{2010CQGra..27q3001A} in considering a single detector with an
SNR threshold $\rho \ge 8$ as a proxy for detectability by the network.
With this criterion, a simple and common expression to transform the local
merger rate to a predicted detection rate $R_D$, given the horizon distance
$D_h$ and the merger rate density, ${\cal R}(z)$, evaluated locally
(at $z=0$), is:
\begin{equation}
\label{eq:LocalUniverseMergerRateFormula}
R_D \simeq \frac{4\pi}{3} D_{h}^3 \langle w^3\rangle \left<({{\cal M}_c}/1.2 M_\odot)^{15/6}\right>{\cal R}(0)
\end{equation}
In this expression $\langle w^3\rangle^{-1/3}\simeq 2.264$ is a
purely geometrical and SNR-threshold-independent factor commonly used to relate sky
location- and orientation-averaged distances to optimal detection distances (see Appendix
for details) and ${{\cal M}_c}=\eta^{3/5}M$ (where $M=m_1+m_2$ is the total mass of
the binary and $\eta\equiv m_1m_2/M^2$) is the ``chirp mass'' \citep[see, e.g.,][]{cutlerflanagan}.
This estimate assumes that (1) cosmological effects are negligible
(i.e., space is Euclidean to a good approximation), and (2) most of
the SNR is accumulated during an inspiral phase which lasts through
the entire sensitive band of the detector, where the GW amplitude in
the frequency domain is well approximated by the quadrupole formula,
i.e., $\tilde h(f)\sim {{\cal M}_c}^{5/6} f^{-7/6}/D$. Here $D$ is the
luminosity distance to the source. The estimate of
Eq.~(\ref{eq:LocalUniverseMergerRateFormula}) follows from this simple
scaling together with the definition of the SNR, Eq.~(\ref{SNR}).
\begin{figure}
\includegraphics[width=1.0\columnwidth,clip=true]{noise}
\caption{\label{fig:noise} \textbf{Noise models}: we use an analytical approximation to the aLIGO zero-detuning
high power (ZDHP) noise power spectral density given in Eq.~(4.7) of \cite{ajithspin}
(we verified that this approximation gives results in excellent
agreement with the ``official'' tabulated aLIGO ZDHP noise
PSD given in \cite{PSD:AL}. For AdV we use the fit in Eq.~(3.4) of
\cite{ajithbose} to \cite{AdvVirgo}, and for KAGRA we use the PSD fit from the Appendix of
\cite{pannarale} to \cite{KAGRA}.}
\end{figure}
Eq.~(\ref{eq:LocalUniverseMergerRateFormula}) involves only the
\emph{local} merger rate ${\cal R}(0)$ and $\langle {{\cal M}_c}^{15/6}\rangle$
is averaged over detected binaries. Both quantities can easily be
extracted from {\tt StarTrack} simulations; they are listed in Table
\ref{tab:simplerates}, along with the values of $R_D$ predicted by
Eq.~(\ref{eq:LocalUniverseMergerRateFormula}). We expect this rough
estimate to be accurate for NS-NS binaries, for which the overwhelming
majority of the SNR is accumulated during the inspiral phase. More
accurate calculations are required for DCOs comprised of BHs, because
they are visible out to larger distances (making cosmological
corrections important) and because, as we discuss below, a large
fraction of the SNR for these binaries comes from the merger/ringdown
portion of the signal.
\subsection{Including merger and ringdown} \label{sec:IMR}
In order to refine our rate estimates for high-mass systems containing
BHs, it is important to consider the full waveform, including
inspiral, merger, and ringdown (IMR). The calculation of gravitational
waveforms from merging BH-BH and BH-NS binaries requires expensive
numerical relativity simulations, but several semi-analytical models
have been tuned to reproduce the amplitude and phasing of BH-BH and
BH-NS merger simulations. To estimate systematic uncertainties and the
impact of spin, we performed rate calculations using three models: (1)
the IMRPhenomB model described in \cite{PhenomB}, one of the earliest
phenomenological models tuned to both nonspinning and spinning BH-BH
simulations with aligned spins, henceforth abbreviated as PhB; (2) the
IMRPhenomC (henceforth abbreviated PhC) model by \cite{santamaria}, a
more accurate alternative to PhB also tuned to nonprecessing
simulations of BH-BH mergers; and (3) a nonspinning effective-one-body
(EOB) model \citep{eob}. A detailed comparison of the three models can
be found in \cite{Damour:2010zb}.
Recent work by \cite{pannarale} shows that finite-size effects
introduce negligible errors ($\lesssim 1\%$) in SNR calculations for
BH-NS binaries, therefore the above models are adequate for {\em both} BH-BH
and BH-NS binaries.
In order to facilitate comparison with previous work, we also
evaluated rates using the simplest possible approximation: a
restricted post-Newtonian (PN) waveform where the amplitude is
truncated at Newtonian order, i.e. $\tilde h(f)\sim {{\cal M}_c}^{5/6}
f^{-7/6}/D$, terminated at a fiducial ``innermost stable circular
orbit'' frequency $f_{\rm ISCO}=(G M\pi/c^3)^{-1}6^{-3/2}$. At low
mass, the upper limit can be neglected and this approximation
corresponds to $\rho \propto {{\cal M}_c}^{5/6}$, as stated above: see also
Eq.~(7) in \cite{roskb}.
\begin{figure}
\includegraphics[width=\columnwidth]{fig-mma-paper-SNRVersusMass}
\caption{\label{fig:Ingredients:SNRVersusMass:CompareModels}\textbf{SNR
for different signal models}: To illustrate the relatively small
differences between the signal models we have adopted, we show the
SNR, $\rho(M)$, as a function of total binary mass, $M$, for an
equal-mass nonspinning binary at $100 \unit{Mpc}$, where the SNR is
evaluated using a single fiducial aLIGO detector. The colored solid
curves show (a) the trivial expression $\rho =\rho_0(M/2.8
M_\odot)^{5/6}$ with $\rho_0=34.3$ (red), (b) an EOB model (black),
PhB model (blue), and PhC model (green), all evaluated for zero
spin. The green dotted line shows the PhC model evaluated with
near-extremal spin on both objects ($\chi_1=\chi_2=0.998$), while
the green dashed line shows PhC with near-extremal spin on one
object ($\chi_1=0.998,\chi_2=0$). The choice $\chi_i=0.998$
corresponds to the \cite{Thorne:1974ve} bound. This value of the
spin is outside the regime in which phenomenological models have
been calibrated, and it has been chosen to provide rough upper
limits on the rates.}
\end{figure}
Figure \ref{fig:Ingredients:SNRVersusMass:CompareModels} shows that
these models all make similar predictions for the SNR of optimally
oriented equal-mass binaries as a function of their total mass for a
single aLIGO detector. Even small differences can be important: for
any given binary, a $30\%$ difference in amplitude corresponds to a
factor $(1.3)^3\simeq 2.2$ in rate calculations. In practice, however,
all nonspinning IMR models agree in SNR to within tens of percent over
the total binary mass range of interest (up to $127{\rm ~M}_{\odot}$, see
Section~\ref{dcos}). The effect of spin will be discussed in more
detail in Section~\ref{subsec:WF} below.
\begin{figure*}
\ifpdf{
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_SNR100Mpc_ins}
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_SNR100Mpc_IMR}
}
\else
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_SNR100Mpc_ins.ps}
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_SNR100Mpc_IMR.ps}
\fi
\caption{\label{fig:Ingredients:SNRContours}\textbf{Optimal SNR for
nonspinning binaries} of given (redshifted) total mass $M_z=M(1+z)$
and mass ratio $q=m_2/m_1$ at luminosity distance $D_L=100$~Mpc. In
the left panel the SNR is computed using the restricted PN
approximation (i.e., the GW amplitude is evaluated using the
quadrupole formula). In the right panel we use the PhC model for the
full IMR waveform; the results for the EOB model are very similar.
A low-frequency cutoff of $f_{\rm cut}=20$Hz has been assumed (see
Section \ref{subsec:WF} for further details).}
\end{figure*}
\begin{figure*}
\ifpdf{
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_DLhor_ins}
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_DLhor_IMR}\\
}
\else
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_DLhor_ins.ps}
\includegraphics[width=1.0\columnwidth,clip=true]{AdLIGOZDHP_20Hz_DLhor_IMR.ps}\\
\fi
\caption{\label{fig:Ingredients:Dhorizon}\textbf{Horizon luminosity distance
(in Mpc) for nonspinning binaries} as a function of redshifted mass
and mass ratio, computed according to Eq.~(\ref{DLhor}) using
waveforms comprised of only the inspiral (left panel) or the full
IMR signal (right panel). }
\end{figure*}
Figure \ref{fig:Ingredients:SNRContours} shows contour plots of the
SNR, $\rho$, in the $(M_z, q)$ plane, where $M_z \equiv M(1+z)$ is the
redshifted total mass, $z$ is the redshift, and $q=m_2/m_1 \leq 1$ is
the mass ratio of the components, for nonspinning binaries at
luminosity distance $D_L=100$~Mpc. We discuss the justification for
considering the SNR as a function of $M_z$ below, but since the chosen
distance corresponds to a negligible redshift $z\simeq 0.023$ using
the cosmological parameters $\Omega_M=0.3$, $\Omega_\Lambda=0.7$,
$\Omega_{\rm k}=0$, and $h=0.7$ (chosen for consistency with
\cite{dominik,dominik2}), $M\simeq M_z$ at this distance. The left
panel refers to a calculation using an inspiral-only waveform with
Newtonian amplitude to compute the horizon distance. The right panel
includes inspiral, merger, and ringdown, modelled using the PhC
waveform. This plot shows two important features: (1) including the
full IMR increases the maximum SNR at this luminosity distance by
factors of a few with respect to an inspiral-only calculation, from
$\approx 300$ to $\approx 10^3$; (2) high-mass binaries ($M_z \gtrsim
10^{2.5}M_\odot \approx 300 M_\odot$) involving BHs that would not be
detectable using inspiral waveforms, become detectable using IMR
waveforms. The latter point is not important for the field binaries
considered in this paper, but it is crucial for intermediate-mass BH
mergers \citep[e.g.,][]{walczak,Fregeau:2006,Amaro:2006imbh}.
In an expanding Universe, GW emission is redshifted by the same factor
of $(1+z)$ as electromagnetic radiation. In the units ($G=c=1$)
adopted by relativists to describe gravitational waves, the only
quantity with dimensions in the GW signal is the total mass $M$. Since
the total mass sets the time scale, a binary source of mass $M$ in the
local universe has an identical waveform (but with different
amplitude) to a binary at redshift $z$ with mass $M/(1+z)$; see, e.g.,
\cite{1998PhRvD..57.4535F}. Eq.~(\ref{SNR}), together with the fact
that gravitational amplitudes scale inversely with the luminosity
distance $D_L(z)$, implies that the horizon redshift $z_{\rm h}$
(i.e., the redshift at which an optimally located and oriented binary
would have SNR $\rho_{\rm thr}=8$) can be found via the simple scaling
\begin{equation}\label{DLhor}
D_h (z_{\rm h}) = D_L(z) \frac{\rho(M_z, q)} {\rho_{\rm thr}}\,,
\end{equation}
where $\rho$ is the SNR at any redshift $z$, or luminosity distance
$D_L(z)$. Note that the right-hand side depends only on $z$, $M_z$
and $q$. Therefore one can easily turn an SNR calculation at fixed $z$
(cf. Fig.~\ref{fig:Ingredients:SNRContours}) into a plot of the
horizon luminosity distance $D_h$ (or equivalently of the horizon
redshift $z_h$) such as Fig.~\ref{fig:Ingredients:Dhorizon}.
{\tt StarTrack} produces large catalogs of DCOs with intrinsic
parameters $(M,q)$, with each of these binaries merging at a different
redshift. Any of these representative DCOs is potentially detectable
(depending on precise sky location and binary orientation) when
$z<z_{\rm h}$. Determining detectability therefore amounts to a simple
interpolation of two-dimensional grids similar to those plotted in
Fig.~\ref{fig:Ingredients:Dhorizon}. These grids can be computed once
and for all, given a waveform model and a detector's power spectral
density (PSD). Evaluating such a grid typically involves $100\times
100=10^4$ SNR evaluations, and it is much faster than the
(impractical) evaluation of millions of SNR integrals such as
Eq.~(\ref{SNR}), one for each representative binary produced by {\tt
StarTrack}. The conversion between detectability at optimal location
and optimal orientation and detectability at generic orientations
involves a simple geometrical factor $p_{\rm det}$, as discussed
below.
\section{Rate calculation} \label{sec:fullrates}
The detection rate is
\begin{eqnarray} \label{Rdet}
R_{\rm det} = \iiint_0^\infty {\cal R} (z_m) \frac{dt_m}{dt_{\rm det}} p_{\rm det} \frac{dV_c}{dz_m} dz_m dm_1 dm_2,
\end{eqnarray}
where ${\cal R} (z_m) \equiv \frac{dN}{dm_1 dm_2 dV_c dt_m}$ is the binary merger rate per
unit component mass per unit comoving volume $V_c$ per unit time $t_m$ as measured in the
source frame at merger redshift $z_m$, the term
$\frac{dt_m}{dt_{\rm det}} = \frac{1}{1+z_m}$ accounts for the difference in clock rates
at the merger and at the detector, and $p_{\rm det}=p_{\rm det} (z_m; m_1, m_2)$ is the probability (over isotropic sky locations and orientations) that a source with given masses at a given redshift will be detectable. The quantity
\begin{equation}
\frac{dV_c}{dz}=\frac{4\pi c}{H_0}\frac{D_c^2(z)}{E(z)}\,,
\end{equation}
with $E(z)=\sqrt{\Omega_{\rm M}(1+z)^3+\Omega_{\rm k}(1+z)^2+\Omega_{\Lambda}}$,
is the comoving volume per unit redshift, and
\begin{equation}
D_c(z)=\frac{c}{H_0}\int_0^z \frac{dz'}{E(z')}
\end{equation}
is the comoving distance, related to the luminosity distance $D_L(z)$
by $D_c(z)=D_L(z)/(1+z)$: see \cite{hogg} for our notation and
conventions.
The merger rate ${\cal R} (z_m)$ is a convolution of the star formation rate and the number
density of binaries per unit star forming mass $M_f$ per unit time delay between formation and merger $\tau$:
\begin{eqnarray} \label{Rzm}
{\cal R} (z_m) &=& \int_0^{t_m} \int_0^{t_{\rm det}} \frac{dM_f}{dV_c dt_f} (z_f)
\frac{dN} {dM_f dm_1 dm_2 d\tau} (t_f; m_1, m_2, \tau) \nonumber\\
&\times& \delta(t_m-t_f-\tau) d\tau dt_f,
\end{eqnarray}
where ${\rm SFR} = \frac{dM_f}{dV_c dt_f} (z_f)$ is the star formation rate
per unit comoving volume per unit time $t_f$ at formation redshift $z_f$.
The distribution of binaries in mass and time delay space, $\frac{dN} {dM_f dm_1
dm_2 d\tau} (t_f; m_1, m_2, \tau)$, is obtained with the {\tt StarTrack}
population synthesis code, taking into account the metallicity distribution at
the formation redshift as described in Section~\ref{binevol}. Since {\tt StarTrack} simulations
produce a set of merging binaries with specific component masses and time delays
sampling the desired distribution, the integrals above are easily computed
via Monte Carlo over the simulated systems. For computational efficiency
the outer integral over the time of formation in Eq.~(\ref{Rzm}) is binned over
$\Delta t_f = 100$ Myr segments, while the integral over the merger redshift
$z_m$ in Eq.~(\ref{Rdet}) is transformed into an integral over merger time via $dz_m
= \frac{dz_m}{dt_m} dt_m = H_0 (1+z)E(z) dt_m$ \citep{hogg}.
Thus the detection rate integral can be represented as a Monte Carlo sum over all simulated binaries:
\begin{equation} \label{eq:rate}
R_{\rm det}=\sum \frac{\rm SFR}{\Delta M_f} p_{\rm det} \frac{1}{1+z} \frac{dV_c}{dz} \frac{dz}{dt} \Delta t \,,
\end{equation}
where $\Delta M_f$ is the total star-forming mass that was simulated
in the Monte Carlo to represent the time bin $\Delta t$, all terms but
the first are computed at the merger redshift of the simulated source.
The detection probability for a given source at its merger redshift
$p_{\rm det} (z, m_1, m_2)$ is simply the fraction of sources of a
given mass located at the given redshift that exceed the detectability
threshold in SNR, assuming that sources are uniformly distributed in
sky location and orbital orientation. If a single detector with an
SNR threshold (e.g., $\rho_{\rm thr}=8$) is used as a proxy for
detectability, the detection probability can be expressed as a
cumulative distribution function on the projection parameter $w$. In the Appendix, $w$ is defined such that $w=0$ when the detector has no response to the gravitational wave, and $w=1$ for an optimally located and oriented (face-on and directly overhead) binary. The detection probability is
\begin{equation}
p_{\rm det} = P(\rho_{\rm thr}/\rho_{\rm opt})\,,
\end{equation}
where $P(w)$ is the cumulative distribution function on $w$ over different source locations and orientations, and $\rho_{\rm opt}$ is the signal-to-noise ratio (SNR) for an optimally located
and oriented binary at redshift $z$.
\begin{center}
\begin{deluxetable*}{l|ll|ll}[thb]
\tablewidth{330pt} \tablecaption{Local merger rates and simply-scaled detection rate predictions\tablenotemark{a}:}
\startdata
Model & $\left<{{\cal M}_c}^{15/6}\right>$& ${\cal R}(0)$ & $R_D$ (aLIGO $\rho \ge 8$) & $R_D$ (3-det network $\rho
\ge 10$)\\
& $M_\odot^{15/6}$ & $\unit{Gpc}^{-3} \unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$\\\hline
{NS-NS}\\
Standard & 1.1 (1.1) & 61 (52) & 1.3 (1.1) & 3.2 (2.7)\\
Optimistic CE & 1.2 (1.2) & 162 (137) & 3.9 (3.3) & 9.2 (7.7) \\
Delayed SN & 1.4 (1.4) & 67 (60) & 1.9 (1.7) & 4.5 (4.0) \\
High BH Kicks & 1.1 (1.1) & 57 (52) & 1.2 (1.1) & 3.0 (2.7) \\
\hline
{BH-NS}\\
Standard & 18 (19) & 2.8 (3.0) & 1.0 (1.2) & 2.4 (2.7) \\
Optimistic CE & 17 (16) & 17 (20) & 5.7 (6.5) & 13.8 (15.4) \\
Delayed SN & 24 (20) & 1.0 (2.4) & 0.5 (0.9) & 1.1 (2.3)\\
High BH Kicks & 19 (13) & 0.04 (0.3) & 0.01 (0.08) & 0.04 (0.2)\\
\hline
{BH-BH} \\
Standard & 402 (595) & 28 (36) & 227 (427) & 540 (1017)\\
Optimistic CE & 311 (359) & 109 (221) & 676 (1585) & 1610 (3773)\\
Delayed SN & 829 (814) & 14 (24) & 232 (394) & 552 (938)\\
High Kick & 2159 (3413) & 0.5 (0.5) & 22 (34) & 51 (81)
\enddata
\label{tab:simplerates}
\tablenotetext{a}{Detection rates computed using the basic scaling of
Eq.~(\ref{eq:LocalUniverseMergerRateFormula}) for both the
\textit{high-end} and \textit{low-end} (the latter in parentheses) metallicity
scenarios (see Section \ref{sec:metallicity}). These rates should be
compared with those from more careful calculations presented in
Tables~\ref{rates2genH} and~\ref{rates2genL}.
}
\end{deluxetable*}
\end{center}
\begin{deluxetable*}{l|ll|ll|lll|ll}[thb]
\tablecaption{Detection rates for second-generation detectors in the
\textit{high-end} metallicity scenario}
\startdata
& \multicolumn{2}{c}{AdV [$\rho \ge 8$]}
& \multicolumn{2}{c}{KAGRA [$\rho \ge 8$]}
& \multicolumn{3}{c}{aLIGO [$\rho \ge 8$]}
& \multicolumn{2}{c}{3-det network [$\rho \ge 10 (12)$]}\\
& \multicolumn{2}{c}{$f_{\rm cut}=20$~Hz}
& \multicolumn{2}{c}{$f_{\rm cut}=10$~Hz}
& \multicolumn{3}{c}{$f_{\rm cut}=20$~Hz}
& \multicolumn{2}{c}{$f_{\rm cut}=20$~Hz}\\
Model &Insp &PhC (EOB) &Insp &PhC (EOB) &Insp &PhC (EOB) & PhC (spin) & Insp & PhC\\
& $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$
& $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ \\\hline
{NS-NS}\\
Standard & 0.3 & 0.3 & 0.8 & 0.7 & 1.2 & 1.1 &- & 2.5 (1.5) & 2.4 (1.4) \\
Optimistic CE & 0.9 & 0.9 & 2.1 & 1.9 & 3.3 & 3.1 &- & 6.9 (4.0) & 6.5 (3.8) \\
Delayed SN & 0.4 & 0.4 & 1.0 & 0.9 & 1.6 & 1.5 &- & 3.3 (1.9) & 3.1 (1.8) \\
High BH Kicks & 0.3 & 0.3 & 0.7 & 0.7 & 1.1 & 1.1 &- & 2.3 (1.4) & 2.2 (1.3) \\
\hline
{BH-NS}\\
Standard & 0.2 & 0.2 & 0.5 & 0.4 & 0.7 & 0.6 & 0.8 & 1.5 (0.9) & 1.2 (0.7) \\
Optimistic CE & 1.1 & 1.0 & 2.9 & 2.2 & 4.4 & 3.6 & 4.4 & 9.2 (5.4) & 7.4 (4.3) \\
Delayed SN & 0.09 & 0.07 & 0.2 & 0.2 & 0.4 & 0.3 & 0.5 & 0.8 (0.5) & 0.6 (0.3) \\
High BH Kicks & 0.01 & 0.007 & 0.02 & 0.02 & 0.04 & 0.03 & 0.1 & 0.09 (0.05) & 0.07 (0.04) \\
\hline
{BH-BH} \\
Standard & 35 & 41 (38) & 70 & 93 (86) & 117 & 148 (142) & 348 & 236 (144) & 306 (177) \\
Optimistic CE & 126 & 144 (133) & 281 & 366 (333) & 491 & 618 (585) & 1554 & 1042 (588) & 1338 (713) \\
Delayed SN & 27 & 34 (32) & 50 & 81 (75) & 90 & 129 (124) & 320 & 182 (110) & 270 (155) \\
High Kick & 0.6 & 1.0 (0.9) & 0.9 & 2.5 (2.3) & 2.1 & 3.8 (3.8) & 12 & 4.2 (2.7) & 8.2 (4.7)
\enddata
\label{rates2genH}
\tablenotetext{a}{Detection rates computed for the high-end
metallicity evolution scenario using the inspiral (``Insp'') and PhC
or EOB IMR models for nonspinning binaries. For aLIGO we also list
rough upper limits on the rates computed with the IMR PhC model by
assuming that BHs have near-maximal aligned spins
($\chi_1=\chi_2=0.998$ for BH-BH systems; $\chi_1=0.998$ and
$\chi_2=0$ for BH-NS systems). The inspiral is calculated using the
restricted PN approximation, which overestimates the amplitude (and
therefore the detection rates) for low-mass systems (NS-NS) when
compared to the full IMR calculations; cf.~Section~\ref{wmodels} for
details. The last two columns were computed assuming a minimum {\it
network} SNR of 10 (or 12, in parentheses) for a three-detector
network composed of three instruments located at the LIGO Hanford,
LIGO Livingston, and Virgo sites, all with aLIGO sensitivity. For
each detector, $f_{\rm cut}$ is the assumed low-frequency cutoff in
the power spectral density: see section \ref{subsec:WF}.}
\end{deluxetable*}
\begin{deluxetable*}{l|ll|ll|lll|ll}[thb]
\tablecaption{Detection rates for second-generation detectors in the
\textit{low-end} metallicity scenario}
\startdata
& \multicolumn{2}{c}{AdV [$\rho \ge 8$]}
& \multicolumn{2}{c}{KAGRA [$\rho \ge 8$]}
& \multicolumn{3}{c}{aLIGO [$\rho \ge 8$]}
& \multicolumn{2}{c}{3-det network [$\rho \ge 10 (12)$]}\\
& \multicolumn{2}{c}{$f_{\rm cut}=20$~Hz}
& \multicolumn{2}{c}{$f_{\rm cut}=10$~Hz}
& \multicolumn{3}{c}{$f_{\rm cut}=20$~Hz}
& \multicolumn{2}{c}{$f_{\rm cut}=20$~Hz}\\
Model &Insp &PhC (EOB) &Insp &PhC (EOB) &Insp &PhC (EOB) & PhC (spin) & Insp & PhC \\
& $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$
& $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ & $\unit{yr}^{-1}$ \\\hline
{NS-NS}\\
Standard & 0.3 & 0.3 & 0.7 & 0.6 & 1.1 & 1.0 &- & 2.3 (1.3) & 2.2 (1.3) \\
Optimistic CE & 0.8 & 0.7 & 1.8 & 1.7 & 2.9 & 2.7 &- & 6.0 (3.5) & 5.6 (3.3) \\
Delayed SN & 0.4 & 0.4 & 1.0 & 0.9 & 1.5 & 1.4 &- & 3.2 (1.8) & 2.9 (1.7) \\
High BH Kicks & 0.3 & 0.3 & 0.7 & 0.6 & 1.0 & 1.0 &- & 2.1 (1.3) & 2.0 (1.2) \\
\hline
{BH-NS}\\
Standard & 0.3 & 0.2 & 0.7 & 0.5 & 1.1 & 0.8 & 1.2 & 2.3 (1.3) & 1.8 (1.0) \\
Optimistic CE & 1.4 & 1.2 & 3.6 & 2.8 & 5.5 & 4.4 & 5.7 & 12 (6.7) & 9.4 (5.4) \\
Delayed SN & 0.2 & 0.1 & 0.5 & 0.4 & 0.8 & 0.6 & 0.9 & 1.7 (0.9) & 1.3 (0.7) \\
High BH Kicks & 0.04 & 0.03 & 0.09 & 0.07 & 0.1 & 0.1 & 0.3 & 0.6 (0.2) & 0.5 (0.2)\\
\hline
{BH-BH} \\
Standard & 56 & 66 (61) & 106 & 153 (140) & 183 & 246 (235) & 610 & 369 (226) & 514 (292) \\
Optimistic CE & 287 & 324 (297) & 629 & 828 (745) & 1124 & 1421 (1339) & 3560 & 2384 (1336) & 3087 (1633) \\
Delayed SN & 53 & 64 (59) & 97 & 152 (139) & 171 & 241 (231) & 596 & 345 (213) & 501 (291) \\
High Kick & 0.9 & 1.5 (1.4) & 1.4 & 3.8 (3.6) & 3.2 & 5.9 (5.8) & 19 & 6.6 (4.0) & 13 (7.2)
\enddata
\label{rates2genL}
\tablenotetext{a}{Same as Table~\ref{rates2genH}, but for the
\textit{low-end} metallicity scenario.}
\end{deluxetable*}
\section{Results} \label{sec:results}
In Section~\ref{sec:simplerates}
we obtained a rough estimate of event rates by extrapolating the local
rate density via the scaling of
Eq.~(\ref{eq:LocalUniverseMergerRateFormula}).
This extrapolation is expected to
provide a good approximation for low-mass systems (and in particular,
NS-NS binaries), because in this case the early inspiral makes up
most of the signal observable by advanced GW detectors, the signal extends through the detector band, and the detector range is sufficiently low that cosmological corrections to detectability and the dependence of merger rates on redshift can largely be ignored. The approximation will
become increasingly inaccurate for high-mass binaries, such as those
comprising one or two BHs. In Sections~\ref{sec:IMR} and~\ref{sec:fullrates}
we went beyond this
approximation by implementing three ``complete'' IMR waveform models
(EOB, PhC, PhB), and we described how to combine these models with
simulations from the {\tt StarTrack} code in order to obtain more
accurate estimates of the event rates (see Eq.~(\ref{eq:rate})).
The analytical estimates of Section~\ref{sec:simplerates} with local
merger rates based on the {\tt StarTrack} code are presented in
Table~\ref{tab:simplerates}. The more careful event rate calculations
of Section~\ref{sec:fullrates} are listed in Table \ref{rates2genH}
(for the high-end metallicity scenario) and Table \ref{rates2genL}
(for the low-end metallicity scenario).
In these tables, the ``single-detector'' columns represent estimated detection rates
for a single detector with a $\rho \ge 8$ threshold for detectability. This is often used
as a proxy for rates in multi-detector networks \citep{2010CQGra..27q3001A}. In the
``three-detector'' columns we consider two alternate detectability thresholds: minimum
{\it network} SNRs of either 10 or 12 for a three-detector network composed of three instruments
located at the LIGO Hanford, LIGO Livingston, and Virgo sites, all with aLIGO sensitivity. The
network SNR threshold of 10 would have yielded false alarm rates of roughly once per decade in
2009-2010 initial LIGO and Virgo data \citep[see Fig.~3 in][]{scenarios}. This
threshold is optimistic for making confident detections if data quality in advanced detectors is
similar to that in the initial detectors and the same searches are used. With this in mind,
\citet{scenarios} assumed a network SNR threshold of $12$ with an additional threshold constraint
on the SNR in the second-loudest instrument; we consider a simple SNR threshold of $12$.
Detection rates using a network SNR threshold were calculated using the same
framework as above, but implementing a network-geometry-dependent
$P(w)$ described (and fitted) in the Appendix.
In the order-of-magnitude estimates described by
Eq. (\ref{eq:LocalUniverseMergerRateFormula}) and provided in Table
\ref{tab:simplerates} we employ $\left<w^{3}\right> \simeq 0.404$ for
the three-detector network ($\rho \ge 10$), a factor of $\sim
4.6$ larger than the value $\left<w^{3}\right> \simeq (1/2.26)^3
\approx 0.0866$ used for a one-detector network.
We now discuss these rate predictions, their dependence on
gravitational waveform models, and the astrophysical properties of DCO
populations observable by advanced GW detectors.
\subsection{Broad features of rate estimates}
The main conclusion of this work is
that BH-BH mergers should yield the highest detection rates in all
advanced detectors (aLIGO, AdV, and KAGRA), followed by NS-NS
mergers, with BH-NS mergers being the rarest. This finding is independent of
our evolutionary models and of the details of the gravitational waveforms
(however, see Sec.~\ref{sec:conclusions} for discussion).
The only exception is the ``Optimistic CE'' model, where detection rates for
BH-NS mergers dominate over NS-NS mergers (with BH-BH mergers still
dominating the detection rates). This model makes the
assumption that CE events with HG donors do not always end in
a premature merger, allowing more binaries to survive the CE and form
merging DCOs, and therefore increasing detection rates. As a result
the Optimistic CE model yields very large BH-BH rates, comparable to,
though a factor of a few below, existing upper limits on the BH-BH
binary mergers from initial LIGO/Virgo observations \citep[see,
e.g.,][]{comparison,2012PhRvD..85h2002A,2013PhRvD..87b2002A}.
Our quantitative predictions for compact binary merger rates are consistent
with our previous papers in this series \citep{dominik,dominik2}. In
particular, we agree with the main conclusion of those papers: detectable
BH-BH binaries can be formed over a broad range of metallicities, with a
significant proportion forming in highly subsolar environments
(Fig.~\ref{fig:FiducialResultDistributions:BHBH}).
On a model-by-model basis our results are in good
agreement with prior work, with factor-of-two or smaller differences due to
our inclusion of cosmological effects.
As expected, the simple approximation of
Eq.~(\ref{eq:LocalUniverseMergerRateFormula}) gives a good
order-of-magnitude estimate of the NS-NS detection rates listed in
Tables \ref{rates2genH} and \ref{rates2genL}. However, the
approximation fails for BH-BH systems. By comparing the detection
rates from Table \ref{tab:simplerates} with inspiral rates from Tables
\ref{rates2genH} and \ref{rates2genL}, we see that the local Universe
approximation of Eq.~(\ref{eq:LocalUniverseMergerRateFormula})
overestimates more careful calculations of detection rates by a factor
$\sim 2$ for BH-BH systems. The limited signal bandwidth of high-mass
systems, the redshift dependence of binary merger rates, and
cosmological corrections make simple scaling relations inaccurate over
the large volume in which detectors are sensitive to BH-BH systems.
On the other hand, as the merger--ringdown phase of these binaries
falls within the sensitive band of second-generation interferometers,
it provides a significant contribution to the SNR. Indeed, as can be
seen in Tables \ref{rates2genH} and \ref{rates2genL}, the full IMR
calculations increase the detection rates considerably. However,
BH-BH detection rates computed with appropriate cosmological
corrections are still lower than local merger rates converted into
detection rates via the basic scaling of
Eq.~(\ref{eq:LocalUniverseMergerRateFormula}).
\begin{figure}
\includegraphics[width=0.86\columnwidth]{fig-mma-dPdtb-NSNShigh}
\includegraphics[width=0.86\columnwidth]{fig-mma-logZ-NSNShigh}
\includegraphics[width=0.86\columnwidth]{fig-mma-logdPdMc-NSNShigh}
\includegraphics[width=0.86\columnwidth]{fig-mma-PMc-NSNShigh}
\caption{ \label{fig:FiducialResultDistributions:NSNS}\textbf{Compact
NS-NS binaries detectable by aLIGO}: Properties of NS-NS binaries
with $\rho \ge 8$ in a single aLIGO instrument in the high-end metallicity
scenario, scaled in proportion to their detection probability.
Different color and line styles indicate results for different
binary evolution models: Standard model (solid black), Optimistic CE (dotted black), delayed SN
(dashed black), and high BH kicks (blue). The top, second, and third panels show the
distribution of birth time $t_{\rm f}$, birth metallicity $Z_{\rm b}$ (with a vertical bar
marking solar metallicity, ${\rm Z}_{\odot}=0.02$), and chirp mass
${{\cal M}_c}$, respectively. The bottom panel shows the cumulative distribution in chirp mass,
to highlight significant changes on a linear scale. The time domain ranges from $0$ Gyr (Big
Bang) to $13.47$ Gyr (today). Though our simulations use a discrete
array of metallicity bins, to guide the eye their relative
contributions have been joined by solid lines in the second panel; this histogram makes
no correction for the density of metallicity bins.
}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{fig-mma-dPdtb-BHNShigh}
\includegraphics[width=\columnwidth]{fig-mma-logZ-BHNShigh}
\includegraphics[width=\columnwidth]{fig-mma-logdPdMc-BHNShigh}
\includegraphics[width=\columnwidth]{fig-mma-PMc-BHNShigh}
\caption{\label{fig:FiducialResultDistributions:BHNS} \textbf{Compact
BH-NS binaries detectable by aLIGO}: Same as
Figure~\ref{fig:FiducialResultDistributions:NSNS}, but for BH-NS
binaries in the high-end metallicity scenario. Some of the sharp features in
the chirp mass distribution are an artifact of the crude binning in metallicity
undertaken for computational reasons; see the discussion in section \ref{dcos}.}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{fig-mma-dPdtb-BHBHhigh}
\includegraphics[width=\columnwidth]{fig-mma-logZ-BHBHhigh}
\includegraphics[width=\columnwidth]{fig-mma-logdPdMc-BHBHhigh}
\includegraphics[width=\columnwidth]{fig-mma-PMc-BHBHhigh}
\caption{\label{fig:FiducialResultDistributions:BHBH} \textbf{BH-BH
binaries detectable by aLIGO}: Same as
Figure~\ref{fig:FiducialResultDistributions:NSNS}, but for BH-BH
binaries in the high-end metallicity scenario. Some of the sharp features in
the chirp mass distribution are an artifact of the crude binning in
metallicity undertaken for computational reasons; see the discussion in section \ref{dcos}.}
\end{figure}
\subsection{Impact of waveform models on predicted rates}\label{subsec:WF}
Our results show that the merger-ringdown contribution is not
important for estimating detection rates of DCOs containing NSs. In
fact, when compared with the restricted PN model, the IMR waveforms
slightly {\em decrease} event rates for NS-NS and BH-NS systems.
The reason for this reduction is that IMR waveforms (such as PhC and
EOB) provide a more accurate representation of the early inspiral,
incorporating PN amplitude corrections that {\em reduce} the signal
amplitude\footnote{Note that in
Eq.~(3.14) of \cite{santamaria} the coefficient of the dominant correction, ${\cal A}_2$,
listed in their Eq.~(A5) is negative.}---and hence the event rates---for
signals dominated by the early inspiral.
BH-NS systems may be subject to an additional event rate
reduction mechanism. There is the possibility of the NS being
distorted and disrupted by the BH tidal field. When these violent phenomena
occur, a suppression of the GW amplitude takes place before the ISCO frequency,
and the SNR decreases with respect to that of a BH-BH system with the same
properties. The GW shut--off due to NS tidal disruption depends on the parameters
of the system: large values of the mass ratio, the BH spin, the NS
radius and the low tilt angles of NS orbital angular momentum relative to
BH spin all favor NS disruption (e.g., \cite{kabelspin}).
By using point-particle IMR waveforms to describe the GW
emission of BH-NS systems we are neglecting this event rate reduction mechanism.
While it would be possible to take these effects into account for nonspinning systems by
using the GW amplitude model of \cite{pannarale}, accurate models for systems
with spinning BHs do not exist yet. For consistency we therefore use
BH-BH waveform models in both cases. Additionally, \cite{pannarale} found
that in the nonspinning case, the SNR difference between the mergers of disrupted
BH-NS systems and the undisrupted systems modeled with PhC is less than $1\%$.
Including the merger portion of the signal is important for BH-BH
systems. For illustration, let us focus on the Standard Model: if we
use PhC waveforms rather than the restricted PN approximation, we find a
$\sim25\%$ increase in the detection rates of BH-BH
systems, from 117 (183) to 148 (246) in the
\textit{high-end} (\textit{low-end}) metallicity scenario.
The rates predicted by EOB and PhC models agree quite well\footnote{We
also carried out calculations using PhB models, which overestimate
rates by about $10\%$ with respect to PhC models. We decided not to
report these results in the Tables, because the PhB model is
less accurate than PhC, although it is easier to implement and less
computationally expensive.}. This can be understood by looking again
at Figure \ref{fig:Ingredients:SNRVersusMass:CompareModels}, which
shows that different approximations of the strong-field merger
waveform agree rather well (at least in the equal-mass limit) on the
SNR $\rho$ and hence on the predicted event rates, which scale with
the cube of the SNR. Waveform differences produce systematic rate
uncertainties significantly less than a factor of 2, much smaller than
astrophysical differences between our preferred models.
Our detailed calculation shows that typically PhC models overestimate
the rates by about $10\%$ when compared to EOB models. This agreement
is nontrivial, because the two families of models are very different
in spirit and construction: the PhC family is a frequency-domain model
that can be easily implemented in rate calculations, while the
time-domain EOB model is more accurate in its domain of validity and
more computationally demanding. It is important to note that in order
to use the two families of models in rate calculations we must compute
waveforms and SNRs in regions of the parameter space where the models
were not tuned to numerical relativity simulations. In particular,
both models become less accurate for small mass-ratio binaries.
Besides systematic errors in waveform modeling, the detection rates
reported in this work (and the resulting distribution of detectable
DCO parameters) depend on our detection criteria.
We ignore a variety of complications of the detection pipelines, such
as the difficulty of searching for precessing sources, noise artifacts
(non-stationary, non-Gaussian ``glitches'' in the instruments) which
can make searches for shorter, high-mass signals less sensitive, and
the limited uptime of detectors. Instead, we have assumed several simplistic
detection thresholds on single-detector or network SNR that are constant across all masses and mass ratios.
Moreover, achieving good detector sensitivity at low frequencies may
prove particularly difficult. We have only included bandwidth above
specified low-frequency cutoffs ($f_{\rm cut}=20$~Hz in most cases)
for detection-rate calculations. However, the specific choice of low
frequency cutoff has minimal impact on our results. For example, using
a lower cutoff $f_{\rm cut}=10$~Hz rather than $f_{\rm cut}=20$~Hz in
the single-detector, high-end metallicity aLIGO rate calculation would
increase the Standard Model BH-BH rates from 117 to 128 in the
inspiral case, and from 148 to 161 in the IMR case. The effect is even
smaller for BH-NS and NS-NS rates.
The impact of spins on the predicted detection rates can be important.
We only consider BH spins, since NSs in compact binaries are not
expected to be rapidly spinning
\cite[e.g.,][]{MandelOShaughnessy:2010} and the dynamical impact of NS
spin will be small. In Tables \ref{rates2genH} and \ref{rates2genL} we
use the PhC model to estimate the possible impact of BH spin on BH-NS
and BH-BH detection rates by assuming that all BHs are nearly
maximally spinning (i.e., with dimensionless spin parameter
$\chi_1=\chi_2=0.998$) and aligned with the orbital angular momentum.
Aligned BH spins cause an orbital hang-up effect that increases the
overall power radiated in the merger, produces a rapidly spinning
merger remnant, and therefore increases the range to which high-mass
binaries can be detected.
We find that spin effects may increase BH-BH detection rates by as
much as a factor of $3$. These increased rates are a direct result of
the increased horizon distance to spinning binaries. For example, a
(30+30) $M_\odot$ binary can be observed to roughly $1.3$ times
farther and be detected $\simeq (1.3)^3 \simeq 2$ more often with
near-maximal spins than with zero spin. Additionally, spin dynamics
can provide a direct diagnostic of the dominant physical effects in
DCO formation \citep{gerosa}. Spin effects only marginally increase
BH-NS rates, but (as discussed at the beginning of this section) tidal
disruption, which we neglected, may have the opposite effect.
\subsection{Astrophysical properties of observable DCOs}
\label{dcos}
We now turn to a more detailed analysis of the observable
properties of DCOs. For concreteness we will focus on aLIGO results
for the ``Standard model'' and nonspinning PhC waveforms, unless
stated otherwise.
\smallskip
\noindent
\textbf{NS-NS}.
By comparing Tables \ref{rates2genH} and \ref{rates2genL} we see that
the detection rates of NS-NS systems are not sensitive to our
differing metallicity evolution scenarios. For simplicity, we therefore only
discuss our results for the \textit{high-end} metallicity evolution scenario.
As shown in our previous work \citep{dominik}, NS-NS systems are
efficiently created in metal-rich environments. The observable
population shares this trait, and half of the observable systems
originate from solar metallicities and higher. As the average
metallicity content of the Universe correlates with time and as most DCOs
preferentially merge shortly after formation
(i.e., the time delay distribution is $\propto t_{\rm merger}^{-1}$; see \cite{dominik}),
the birth rate of detectable NS-NS systems peaks at $13$ Gyrs after the Big Bang
(see Fig.~\ref{fig:FiducialResultDistributions:NSNS}). The most
distant detectable system has a merger redshift $z\sim 0.13$ (or luminosity distance
$L_{\rm D}=610$ Mpc).
The range of possible chirp masses in the third panel from the top of
Figure \ref{fig:FiducialResultDistributions:NSNS} is limited at the low end ($>0.87 M_\odot$)
by the $1M_\odot$ minimum birth mass for NS and is limited at the high
end by the (assumed) maximum mass for a NS ($m_{NS} <2.5 M_\odot; {{\cal M}_c} < 2.1 M_\odot$). The birth mass,
in turn, is set by supernova physics, which we have implemented as the Rapid or Delayed SN engine
\citep{chrisija}. For this reason the NS mass difference between the SN engines is intrinsic
to the entire merging population of NS-NS systems. Therefore, this observable feature should be
available to any of the detectors considered in this study.
The chirp mass distributions for Standard and Optimistic CE models span the range from
$0.9{\rm ~M}_{\odot}$ to $1.6{\rm ~M}_{\odot}$. The Delayed SN model results in a notably different NS mass
distribution, favoring heavier masses. As the SN explosion in the Delayed engine lasts longer,
more matter is accreted onto the proto--NS (which is more massive than in the Rapid engine scenario),
allowing the formation of more massive remnants (cf.~Figure~\ref{fig:FiducialResultDistributions:NSNS}).
The maximum allowed NS mass in this model is $2.5{\rm ~M}_{\odot}$, and in extreme
(but very rare) cases this mass is approached; the maximum chirp mass
for a detectable system in our Monte Carlo simulation was $2.1{\rm ~M}_{\odot}$,
corresponding to both components close to the maximum allowed limit. For comparison,
chirp masses of NS-NS systems in the models utilizing the Rapid SN engine
(Standard, Optimistic CE and High BH kick) never exceed $1.7{\rm ~M}_{\odot}$.
Such extremely high masses are rare for all engines, however, and the majority
of chirp masses are much lower, as seen in Figure \ref{fig:FiducialResultDistributions:NSNS}.
The presence of more massive systems in the Delayed SN models extends the horizon
of NS-NS detectability to $z\sim 0.16$ ($L_{\rm D}=765$ Mpc).
Lastly, we note that Standard and High BH kick models are identical for NS-NS systems.
The difference between the black curve (Standard) and blue curve (High BH Kick) in
Figure \ref{fig:FiducialResultDistributions:NSNS} corresponds to the systematic errors
associated with Monte Carlo errors of binary simulations, galaxy sampling, metallicity
binning, etc.
\smallskip
\noindent
\textbf{BH-NS}.
In our previous study \citep{dominik2} we showed that BH-NS
systems are efficiently created at moderate metallicities (${\rm Z}
\sim 0.1\,{\rm Z}_{\odot}$, or $\log({\rm Z})\sim -2.7$). Indeed,
Figure~\ref{fig:FiducialResultDistributions:BHNS} shows that about
half of all detectable BH-NS systems will originate from metallicities
${\rm Z}<0.5\,{\rm Z}_{\odot}$ ($\log({\rm Z})< -2$).
These systems have higher chirp masses than NS-NS systems, on average
$3.3{\rm ~M}_{\odot}$ vs. $1.2{\rm ~M}_{\odot}$, and therefore the detectors can
sample BH-NS systems from a larger volume. However, BH-NS systems are
the rarest of all DCOs per unit (comoving) volume. As a consequence,
BH-NS binaries typically yield the lowest detection rates. One
exception is the Optimistic CE model, in which the merger rate per
unit volume is large enough (while still being lower than for NS-NS
systems at all redshifts) that BH-NS detection rates are larger than
NS-NS rates because they are observed farther (cf. Table~\ref{tab:simplerates} and
Figure~\ref{fig:FiducialResultDistributions:BHNS}).
In our standard model BH-NS systems are detectable up to redshift
$z\approx 0.28$ ($L_{\rm D}=1.4$ Mpc). However, in the Delayed SN model this value reaches
$z\approx 0.31$ ($L_{\rm D}=1.6$ Mpc). As discussed earlier, this is due to the more massive
NSs (up to $2.4{\rm ~M}_{\odot}$) produced by the Delayed engine.
\smallskip
\noindent
\textbf{BH-BH}.
As discussed in our previous papers in this series
\citep{nasza,dominik,dominik2}, BH-BH systems are formed most efficiently in
low-metallicity environments. The detectable population reflects this
property: about half of all detectable BH-BH systems were created in
environments with metallicities ${\rm Z}<0.1\,{\rm Z}_{\odot}$ ($\log({\rm Z})<
-2.7$).
As in prior studies \citep{nasza,dominik,dominik2,vosstauris}, our calculations
imply that BH-BH systems yield the highest detection rates for ground-based
interferometers. This is true even in the ``High BH kick'' model,
where the vast majority of binaries containing a BH are disrupted.
Adjusting the metallicity evolution in the Universe from
\textit{high-end} to \textit{low-end} we see a factor of $\sim 2$
increase in detection rates. In the \textit{low-end}
scenario the average metallicity in the Universe is lower at all times.
Low metallicity environments are much more effective at producing merging
BH-BH systems than higher ones, hence the increase in the detection rates.
Half of the detectable objects have chirp masses above $14{\rm ~M}_{\odot}$.
The most massive of these systems originate from environments with very low metallicity
content (${\rm Z}\sim 0.01\,{\rm Z}_{\odot}$). The birth times of detectable BH-BH systems
peak at $\sim1$~Gyr after the Big Bang. Additionally, half of these systems
were created within $\sim 2$ Gyrs of the Big Bang (see top panel of Figure
\ref{fig:FiducialResultDistributions:BHBH}), when the average
abundance of heavy elements was much smaller than today.
As seen in Tables~\ref{rates2genH} and~\ref{rates2genL}, the detection rates of
BH-BH systems vary as we change our assumptions between the four models and two metallicity
evolution scenarios. By comparing detection rates, for example, found by aLIGO with PhC waveforms,
for the \textit{high-end} metallicity model (works for all model choices), we can distinguish two
extreme configurations: (1) The High BH kick model yields the lowest rates of merging BH-BH systems
($3.8$ yr$^{-1}$). This is a direct consequence of assuming the presence of the maximum natal kick
velocities allowed within our framework, which efficiently disrupt BH progenitor binaries. (2) The
highest detection rate is achieved with the Optimistic CE model ($618$ yr$^{-1}$). Here, it is assumed
that binaries are allowed to progress through the CE with a HG
donor, which adds a significant amount of BH-BH systems to the detectable population.
The detection rates of the other two models: Standard and Delayed SN are similar to each other
($148$ yr$^{-1}$ and $129$ yr$^{-1}$, respectively).
The farthest objects are detectable out to $z\sim2$ ($L_{\rm D}15$ Gpc). These
systems consist of the most massive BH pairs ($m_1=61{\rm ~M}_{\odot}$ and $m_2=66{\rm ~M}_{\odot}$ in the detectable
population, with a chirp mass equal to $55{\rm ~M}_{\odot}$), born $1.8$ Gyr after the Big Bang,
and originating from regions with our lowest considered metallicity content (${\rm Z}=0.005\,{\rm Z}_{\odot}$).
Note that the maximum mass of BH-BH systems is limited by the maximum ZAMS mass of stars,
which was set to $150{\rm ~M}_{\odot}$ in the current simulations. The effect of IMF extending to much
higher masses on detection of BH-BH inspirals have been recently presented by \cite{walczak}.
The detectable BH-BH chirp mass distribution for the Standard model has three
major peaks. These are present at $\sim 7{\rm ~M}_{\odot}$, $14{\rm ~M}_{\odot}$, and $21{\rm ~M}_{\odot}$
(see the black curve in the 3rd and 4th panels of Fig.~\ref{fig:FiducialResultDistributions:BHBH}).
Their presence is associated with the physics governing the Rapid SN engine and the formation
of the most massive BH-BH systems. Within this framework we can distinguish three scenarios
for BH formation, each depending on the pre-SN carbon--oxygen (CO) mass (see Eq.~16 in \cite{chrisija}).
The ``\textit{A}'' scenario occurs for $6{\rm ~M}_{\odot} < M_{\rm CO} \leq 7{\rm ~M}_{\odot}$ and results in full fallback on the BH
and, therefore, no natal kicks (see Eq.~\ref{vkick}). The ``\textit{B}'' scenario occurs for
$7{\rm ~M}_{\odot} < M_{\rm CO} \leq 11{\rm ~M}_{\odot}$, where the fallback is partial and some natal kicks
are present. For this scenario we expect a decreased number of BH-BH systems because of
natal kicks disrupting binary systems during SNe. The ``\textit{C}'' scenario develops for
$M_{\rm CO} \geq 11{\rm ~M}_{\odot}$ and again results in full fallback, and no natal kicks.
BH progenitors originating from ${\rm Z}_{\odot}$ environments never form through the \textit{C}
scenario, since they lose mass in winds at rates that do not allow them to form CO cores larger than
$11{\rm ~M}_{\odot}$. Since BH-BH progenitors in the \textit{B} scenario are subject to disruption due to
the presence of natal kicks, most BH-BH systems in ${\rm Z}_{\odot}$ environments form through the \textit{A}
scenario, with chirp masses clustered around $7{\rm ~M}_{\odot}$.
However, reducing the metallicity by a factor of $2$ lowers the wind mass loss rates
sufficiently to allow BHs to form through the \textit{C} scenario. At this metallicity ($\sim 0.5\,{\rm Z}_{\odot}$)
only the most massive progenitors ($M_{\rm ZAMS}>100{\rm ~M}_{\odot}$) may form BHs through
this scenario. Additionally, the mass of the BHs formed from these high mass components
($M_{\rm ZAMS}>100{\rm ~M}_{\odot}$) only depends weakly on their initial mass. This stems from the fact that
these stars evolve quickly ($\sim\,\mbox{Myrs}$) and lose large fractions of their hydrogen envelope.
Binary evolution does not alter this result significantly, as the interactions between components,
such as mass transfer during CE episodes, also lead to the removal of their hydrogen envelopes.
The result for metallicity $\sim 0.5\,{\rm Z}_{\odot}$ is a clustering of BH-BH systems formed
from the most massive binaries at masses around $16{\rm ~M}_{\odot}$ for each component.
This produces the peak in the chirp mass distribution at $\sim 14{\rm ~M}_{\odot}$.
Reducing the metallicity content by another factor
of $2$ (to $\sim 0.25\,{\rm Z}_{\odot}$) allows the same mechanism to form BH-BH systems with masses
clustering at around $24{\rm ~M}_{\odot}$ for each component. These systems form the peak in the chirp mass
distribution at $\sim 21{\rm ~M}_{\odot}$.
The grouping effect disappears when reducing the metallicity abundance in BH progenitors
even further. For example, at $0.1\,{\rm Z}_{\odot}$ the low wind mass loss rate does not increase the
separation between components as significantly as for higher metallicities. Consequently, the most
massive progenitor binaries engage in a CE phase early in their evolution. This usually happens when
the donor is on the HG and the Standard model does not allow for successful outcomes of such CEs.
However, this scenario is allowed to form BH-BH systems in the Optimistic CE model, yielding the peak
present in the chirp mass distribution at $\sim 29{\rm ~M}_{\odot}$.
As discussed above, the chirp mass distribution in scenario \textit{C}
depends sensitively on the mass loss rate of stars, which depends strongly on metallicity.
Binary evolution for $0.5\,{\rm Z}_{\odot}$ and $0.25\,{\rm Z}_{\odot}$ creates sharp peaks in the chirp mass
distribution of BH-BH systems. In the discrete metallicity grid simulated in this study,
there are no metallicity points between $0.5\,{\rm Z}_{\odot}$ and $0.25\,{\rm Z}_{\odot}$. Targeted follow-up investigations
indicate that metallicity choices between $0.5{\rm Z}_{\odot}$ and $0.25{\rm Z}_{\odot}$ lead to additional sharp peaks in
the chirp mass distribution between $14{\rm ~M}_{\odot}$ -- $21{\rm ~M}_{\odot}$. We expect that an integral over a
fine grid with appropriately small step sizes in metallicity would lead
to all of these narrow peaks merging together to form a single broad distribution without
sharp features. However, we cannot confidently describe the shape of this distribution
without a more detailed investigation with a fine grid of metallicities, which is not
computationally tractable at present.
Finally, the peak in the chirp mass distribution at $\sim 7{\rm ~M}_{\odot}$ in the Standard model is
formed from systems born in $0.5$--$1\,{\rm Z}_{\odot}$ environments. These are low-mass BHs (usually
$8$--$9{\rm ~M}_{\odot}$ per component) formed in the \textit{A} scenario. This formation
is particularly interesting as it does not appear in the Delayed SN
model, with the difference stemming from the different fallback scenarios in the Rapid and Delayed
engines. With the Rapid engine, we can distinguish the three fallback regions.
However, the Delayed engine predicts one region of partial fallback for $3.5{\rm ~M}_{\odot} < M_{\rm CO} \leq 11{\rm ~M}_{\odot}$
and one region of full fallback $M_{\rm CO} \geq 11{\rm ~M}_{\odot}$ (identical to the \textit{C} scenario in the
Rapid engine). Since partial fallback implies the presence of natal kicks and, therefore,
increased probability of binary disruption, there are no ``preferred'' masses for the lightest
BHs in the Delayed SN engine (see dashed line on the 3rd panel,
Fig.~\ref{fig:FiducialResultDistributions:BHBH}) as in the Rapid engine.
The Standard and Delayed SN models also yield different lower mass
limits for BH remnants (see Section \ref{binevol}). For the ``Rapid
engine'' scenario the lowest-mass BH is $\sim 5{\rm ~M}_{\odot}$, while for the
``Delayed engine'' scenario the lowest-mass BH is $\sim 2.5{\rm ~M}_{\odot}$
(this is also the highest NS mass adopted in our {\tt StarTrack}
calculations). As a result, the detectable systems with the lowest
total mass have ${{\cal M}_c}=4.8{\rm ~M}_{\odot}$ and ${{\cal M}_c}=2.4{\rm ~M}_{\odot}$ in the
Rapid and Delayed engine scenarios, respectively.
Additionally, regardless of our evolutionary models the majority BH-BH
systems are formed with nearly equal mass components. Therefore, systems
with mass rations $\sim 1$ dominate the detected population, as shown
in Fig.~\ref{fig:qdis}. For the Delayed SN model the detectable BH-BH systems
with the lowest mass ratio have $q\approx 0.05$. For the remaining models
this value is $q\approx 0.12$.
\begin{figure}
\includegraphics[angle=270,width=\columnwidth]{q.ps}
\caption{\label{fig:qdis} \textbf{Mass ratio ($q$) detection probability distribution for BH-BH
systems.} It is clear that one should expect that the vast majority of detectable BH-BH systems
will be formed of nearly equal mass components. The lowest values of $q$ among the detected systems are
$0.05$ for the Delayed SN model and $0.12$ for the remaining models. For each model the probability
is normalized to the total number of detections for this model.}
\end{figure}
For future reference we also present the initial--final mass
relation for close BH-BH systems in Fig.~\ref{fig:bmr}. The relation is divided into the primary
(more massive at ZAMS) and secondary (less massive) component for two metallicity values
(${\rm Z}_{\odot}$ and $0.1{\rm Z}_{\odot}$), for the Standard model. It is clearly visible that
binary evolution distorts the initial-final mass relation for single stars in both mass dimensions.
In the initial mass dimension, the absence of BHs forming from stars with ZAMS mass above $\sim 70{\rm ~M}_{\odot}$
is a direct consequence of the assumption of the negative (merger) CE outcome for HG donors in our Standard
model. In our framework more massive stars have larger radii and, therefore, are more likely to engage in
CE while the donor is on the HG rather than on later evolutionary stages. If this assumption was relaxed
(Optimistic CE model) the maximum BH mass reached in close BH-BH systems is found to be $150{\rm ~M}_{\odot}$ for both
metallicities. In the final mass dimension, binary evolution prevents remnant components
from reaching masses as high as those formed from single progenitors. Whereas single stars
shed mass only through winds, binaries may also remove mass through interactions like the non-conservative
mass transfer and/or CE events, which consequently lowers the mass of the remnants.
\begin{widetext}
\begin{center}
\begin{figure}
\includegraphics[width=\textwidth]{binary_mass_rel.eps}
\caption{\label{fig:bmr} \textbf{Initial-final mass relation for binary systems.} Presented
for close BH-BH systems, Standard model. We define primary and secondary components as the initially
(at ZAMS) more and less massive, respectively. The shaded scale (right side of each panel) shows the
fractional contribution of a given ZAMS mass bin to the total mass of merging black holes formed
from primaries (left panels) and secondaries (right panels). Note that binary evolution produces a very
different initial-final mass relation than the single stellar evolution (thin line). The top panels and
bottom panels show results for ${\rm Z}_{\odot}$ and $0.1{\rm Z}_{\odot}$, respectively.
}
\end{figure}
\end{center}
\end{widetext}
The initial-final mass relation (in this case for the binary population of
close BH-BH systems) is a result of a number of various initial and
evolutionary assumptions used in population synthesis calculations. Change
of any of these assumptions (whether in initial conditions or evolutionary
calculations) may potentially influence the initial-final mass relation and in
turn the generated BH-BH population. The largest impact is expected from the
treatment of RLOF stability (i.e., criteria for CE development), SN
explosion physics, wind mass loss and internal mixing within massive stars
induced by convection and/or rotation that sets the radial evolution of massive
stars. It seems that the change in the assumptions underlying the initial-final
mass relation may yield no BH-BHs \citep{mennekens} or numerous BH-BH systems
\citep{voss,nasza,dominik,dominik2}. However, these results apply only to isolated binary evolution.
New studies of globular clusters suggest that, such environments may be the birthplaces of
a significant number of BH-BH systems \citep{gcbhbh}.
Note, that the above relations apply only to BH-BH systems. However, our models do not inhibit
the creation of NS from progenitors much more massive than $20 {\rm ~M}_{\odot}$. In fact, the study by
\cite{betaam2008} shows that, due to binary evolution, NS may form from progenitors as massive
as $100 {\rm ~M}_{\odot}$.
\section{Questioning the no BH-BH theorem}\label{sec:nobhbh}
During more than a decade of research into the evolution of binary stars and the formation of
DCOs, several authors proposed the absence of stellar-mass BH-BH systems merging within the Hubble time
(e.g. \cite{nele2001,mennekens}). In the latter study the authors have claimed that the main
reason for this are the high wind mass loss rates experienced by BH progenitors. For example, in
their version of the Brussels population/galactic code (originally \cite{ddv04}) they fix the
wind mass loss rates of the Luminous Blue Variable (LBV) phase at $10^{-3}{\rm ~M}_{\odot}$ yr$^{-1}$.
Following such heavy mass loss, the orbital separation of the components increases
so that they do not engage in CE. As the CE is a major mechanism for reducing orbital separation
in isolated binary evolution, allowing for the formation of close BH-BH systems, the result is an
absence of BH-BH systems detectable through gravitational waves. These results stand in contrast with the
works of \cite{voss} and our previous studies \citep{nasza,dominik,dominik2}.
There are mitigating factors to the finding of \cite{mennekens}. For example, their code does not allow
for tidal interactions between close binary components. As we demonstrate in the
following text, tidal interactions may (even for very high LBV winds) allow
for the formation of close BH-BH binaries (for more on the importance of tidal interactions see e.g.,
\cite{serena}). Let us consider the following example of binary
evolution generated with the StarTrack code. We start with an evolved binary: a $8{\rm ~M}_{\odot}$ BH
accompanied by a $43{\rm ~M}_{\odot}$ companion at the beginning of the HG phase, with an orbital separation of
$4600{\rm ~R}_{\odot}$ at $5.5$ Myr after the creation of the systems (ZAMS). This is a typical phase of a BH-BH progenitor
in our Standard model. In this example we also set the LBV wind mass loss rate to $10^{-3}{\rm ~M}_{\odot}$ yr$^{-1}$
and disable tidal interactions between the components, both as in \cite{mennekens}. We find that
intense wind mass loss widens the orbital separation between the components to such extent that they
never interact. Therefore, when the BH companion forms a second BH, the resulting BH-BH
systems is too wide to merge within a Hubble time. This example is presented in Fig.~\ref{fig:notides}.
\begin{figure}
\includegraphics[angle=270,width=\columnwidth]{notides.ps}
\caption{\label{fig:notides} \textbf{Orbital evolution with tidal interactions disabled}.
This figure presents a part of the evolution of a $8{\rm ~M}_{\odot}$ BH and $43{\rm ~M}_{\odot}$ HG system, with
Luminous Blue Variable wind mass loss rate set at $10^{-3}{\rm ~M}_{\odot}$ yr$^{-1}$.
The top panel shows the evolution of the radius and Roche lobe of the HG star in
addition to the orbital separation in the binary. The bottom panel shows the evolution of the
HG star's spin frequency relative to the orbital frequency. The HG star's activity
as a Luminous Blue Variable is marked by the 'LBV' label. The vertical line separating the 'HG'
and 'CHeB' labels marks the transition of the HG star to the Core Helium Burning phase.
Note that without tidal interactions the binary's orbit expands (due to
stellar wind mass loss) and no component interaction (e.g., CE) is expected.
In the end a wide BH-BH binary is formed.
}
\end{figure}
We can repeat our exercise can be repeated with tidal interactions between the components enabled.
Investigating the same system we find a drastically different outcome of the evolution (see
Fig.~\ref{fig:tides}). As in the example above, the BH companion starts its significant
evolutionary expansion across Hertzsprung gap. Due to the conservation of angular momentum, the expansion
of the star slows its rotation down almost to a standstill.
Once the companion star fills a sizable fraction of its Roche lobe ($\sim 50\%$), the tidal torques
imposed on the star by an orbiting BH transfer the orbital angular momentum into the star, spinning it
up. At first this effect is negligible. However, after approximately $5000$ years, when the radius of
the star becomes sufficient ($\sim 1100{\rm ~R}_{\odot}$), the spin up of the HG star stalls and overpowers the increase of
orbital separation. From this point on, the orbital separation starts to decrease for another
$3000$ years. Finally, when the radius of the star is $\sim 2000{\rm ~R}_{\odot}$, it fills its Roche lobe
and initiates a CE.
\begin{figure}
\includegraphics[angle=270,width=\columnwidth]{tides.ps}
\caption{\label{fig:tides} \textbf{Orbital evolution with tidal interactions enabled}.
Same as Fig.~\ref{fig:notides} but with tidal interactions enabled. The `Rad.~Env.' and
`Conv.~Env.' labels along with corresponding arrows highlight areas where the HG star has a radiative
and convective envelope, respectively. The vertical line linking the arrows marks the transition
point in the structure of the envelope.
Tidal interactions allow the transfer of orbital angular momentum into the
expanding HG star. The associated orbital decay leads to RLOF and the development of a
CE, which allows for the formation of a close BH-BH binary. The timescale on the horizontal axis is
zoomed in relative to Fig.~\ref{fig:notides}.
}
\end{figure}
Our exercise clearly shows that different assumptions may lead to qualitatively different
outcomes in terms of the close BH-BH formation. In particular, assumptions used in this study on
LBV winds, tidal interactions and radial expansion result in a large number of BH-BH mergers.
In contrast, assumptions used by \cite{mennekens} result in no BH-BH mergers formed
out of the isolated binary evolution.
There are several caveats in this framework. First, it is not theoretically well established if stellar
radii can grow to $\sim 2000{\rm ~R}_{\odot}$. For example, intensive mixing (either invoked by rapid rotation or
extended convection in the stellar interior) may reduce the size of the H-rich envelope
which is responsible for expansion in massive stars. On the other hand the
intense wind mass loss may additionally reduce the envelope (e.g., \cite{yusof2013},
but see MESA models for very massive stars \citep{walczak}).
However, the radii of AH Sco, KW Sgr and UY Scuti estimated with
the PHOENIX stellar atmosphere model \citep{wittkowski} extend well beyond $1000{\rm ~R}_{\odot}$, with UY Scuti,
reaching $1708{\rm ~R}_{\odot}$ \citep{yuscuti}. The mass of UY Scuti is estimated to be within
$25{\rm ~M}_{\odot}$--$40{\rm ~M}_{\odot}$, i.e., within the mass range for BH progenitors in our framework. Second, the
efficiency of tidal interactions depends on the structure of the envelope of the participating components.
Stars with convective envelopes tend to respond more strongly to tidal dissipation than stars with radiative
envelopes. In {\tt StarTrack} (see Section 3.3 of \cite{startrack}) we calibrate this phenomenon against
the cutoff period for circularization of a population of MS binaries in M67 and the orbital decay
accompanying tidal synchronization in the LMC X-4 high mass X-ray binary.
This treatment of tidal dissipation applies directly to the given example as the envelope of the
companion star turns from radiative to convective about $3000$ years after the companion enters the
HG (when HG star radius increases to over $\sim 1000 {\rm ~R}_{\odot}$).
However, our simulations show that switching
tidal dissipation to the weaker radiative damping does not prevent binaries from
initiating the CE. In our framework tides are applied to the entire star and
we assume that stars rotate non-differentially. It cannot be excluded that
tides operate only on the outer layers of stellar atmosphere that holds
only a small fraction of a star's mass. Additionally, if there is no (or very
weak) transport of angular momentum within a star, only a
small fraction of orbital energy is used to synchronize the stellar atmosphere
as compared to our prescription. Finally, the moment of inertia of very massive stars depends
strongly on the radial profile, and the {\tt StarTrack} assumptions may yield a moment of inertia
that is too large, therefore providing a more significant reservoir for depositing orbital angular
momentum into the star than is available in practice. If in fact only very little orbital angular
momentum is used for binary component synchronization {\em and} if the winds are
in fact as intense as indicated by \cite{mennekens}, then this would bar the
formation of many close BH-BH binaries found within the framework of our
evolutionary model.
Even if tidal interactions turn out to be ineffective in massive close binaries, this
does not necessarily rule out the formation of close BH-BH binaries. In field populations
about 10--30\% of binaries are, in fact, triples (or higher multiples; e.g., \cite{kiminki1,kiminki2,duchene}) and
Kozai-Lidov effects or dynamical instabilities \citep{PeretsKratter:2012} may lead to the merger
of wide BH-BH binaries.
Additionally, many \citep{kroupa2014} massive stars are formed in clusters and may be subject to
dynamical interactions that can potentially decrease orbital separations.
Finally, over the last few years it has been claimed that dense globular clusters may produce significant
number of close BH-BH binaries. In contrast with earlier findings with no efficient formation of close BH-BH
binaries (e.g., \cite{kulkarni,sigurdsson,zwart2000,banerjee}) the new paradigm emerged based on
recent and updated Monte Carlo simulations of dense cluster evolution (e.g., \cite{mackey,morscher,sippel,
heggie2014}). BH-BH binaries may also form via dynamical interactions in galactic nuclear clusters with or
without a massive black hole \citep{OLeary:2008,MillerLauburg:2008} (but cf.~\cite{Tsang:2013}).
\section{Conclusions}\label{sec:conclusions}
We have calculated cosmological detection rates of merging DCOs for
second-generation GW observatories. We used redshift distributions of
merging DCOs from the {\tt Startrack} population synthesis code, and
have incorporated the cosmic star formation rate as well as galaxy and
metallicity evolution. Using state-of-the-art gravitational waveforms
and detector sensitivity curves, we have translated the cosmological
merger rates into detection rates for four distinct models of binary
evolution.
Our study has several robust implications for imminent GW
searches. First and foremost, our four models agree on the detection rates
of merging NS-NS systems ($\sim 1$ detection per year), with the exception of the Optimistic CE
model which predicts rates a factor of $2$--$3$ times higher than
other models. The mass distributions of detectable NS-NS systems are also similar across the models,
with the exception of the Delayed SN model, which allows for the
formation of NSs with higher masses due to prolonged accretion during
the SN explosion. We predict that NS-NS binaries will be detectable up to redshift
$z\approx 0.13$, i.e., only in the local Universe.
Second, BH-NS systems are expected to be the rarest detectable DCOs (less than $1$ detection
per year), with the exception of the Optimistic CE model, in which BH-NS
detection rates slightly exceed those of NS-NS systems of the same model. We predict
BH-NS systems to be detectable up to redshift $z\approx 0.3$.
In contrast, BH-BH systems will provide the largest number of
detections ($\sim 100$--$1000$ per year), making them the primary target for first detection
and the most promising source for future statistical studies of source
populations. BH-BH systems dominate event rates even in the
pessimistic ``High BH kick'' model (several events per year), wherein most of the systems
containing BHs are disrupted during the SN. Additionally, the BH-BH
mass distribution could have rich, observationally-accessible
structure (various lower limits and shapes) that encodes fine details
about stellar and binary evolution \citep[see,
e.g.,][]{PSconstraints3-MassDistributionMethods-NearbyUniverse,massgap,2012ApJ...757...36K,chrisija}.
We note, however, that the crude binning in metallicity that we had to
undertake in order to limit computational costs may create artificial
sharp, narrow features in the mass distribution, which would merge
together into broader trends with a finer metallicity grid.
\cite{mennekens} point out that
the detection rate of BH-BH systems may be reduced to zero due to the effects of
intense stellar wind during the Red Supergiant and Luminous Blue Variable phases of BH progenitors.
However, we have demonstrated that the \cite{mennekens} result is
a direct consequence of their assumption of no tidal interaction in close
binaries. If tides can efficiently transfer angular momentum from the orbit
into the companion spin, then it is expected that isolated binaries will
form close BH-BH systems.
The criteria for the development of the CE phase may influence the merger and detection rates
of all DCOs. \citep{woods} and \citep{ivanova2014} state that
the criterion for the stability of mass transfer sourced from the polytropic approximation
is much too strict. Therefore, the frequency of the CE may be overestimated. The CE
is a major mechanism for creating close binaries that coalesce within a Hubble time.
The lack of CE events would, therefore, decrease the number of DCO mergers.
This would provide a reasonable pessimistic scenario for the lack of detections
of gravitational wave signals. A study of CE development criteria and its effect on the
formation of close BH-BH binaries is underway (Belczynski et al., in prep.).
However, an assumed rarity of CE systems would be difficult to reconcile
with observational evidence pointing to systems (for example V1309 Sco, V4332 Sgr,
OGLE 2002-BLG-360 or CK Vul) which seem to have developed a CE (e.g., \cite{tylenda,martini99,
tylenda2013}). Additionally, massive X-ray binaries such as NGC300 X-1 or IC10 X-1
are on close orbits with orbital periods $\sim 30$ hr, which have likely developed through a CE event.
Our study shows that detectable NS-NS systems are formed significantly
later in the history of the Universe than BH-BH and BH-NS systems. As
shown in Figs.~\ref{fig:FiducialResultDistributions:NSNS},
\ref{fig:FiducialResultDistributions:BHNS}, and
\ref{fig:FiducialResultDistributions:BHBH}, the birth times of NS-NS
systems cluster around $13$ Gyr after the Big Bang, while for the
other systems this is $1$ Gyr. This behavior might be
counter-intuitive, as the intrinsic distribution of time delays between
formation and merger for all types of DCOs falls off as $t_{\rm merger}^{-1}$,
barring exceptional circumstances \citep[e.g., near-solar metallicity
BH-BH binaries,][]{dominik}. Therefore, one might expect the
majority of detectable DCOs to be formed within the past $\sim$ Gyr as
is the case for NS-NS systems. However, BH-BH systems are created
most efficiently in the lowest metallicity environments, and therefore
their formation rate is highest in the early Universe. The long
time-delay tail of these early systems dominates the subsequent
detection rate. The metallicity evolution is therefore a crucial
factor in predicting the detectable rate of DCOs.
We also find that including the merger and ringdown components of the
GW signal does not have a significant impact on the detection rates of
NS-NS systems. The full IMR calculations become important for higher
mass systems, and especially for BH-BH binaries. The detection rates
for BH-BH systems increases by at least $20\%$, and typically by $\sim
50\%$, when using full IMR waveforms when compared to the PN inspiral
alone.
The detection rate of BH-BH systems is also sensitive to spin
effects. Extreme aligned spins increase the rates by a factor of $\sim
3$ when compared with the non-spinning case.
We used simplified criteria for detectability, considering an SNR
threshold of $8$ in a single detector as a proxy for the network
\citep[cf.~][]{2010CQGra..27q3001A}. For reference, we also
considered a network SNR threshold of $10$, which is likely to be very
optimistic, and $12$, which is more realistic \citep[cf.~][]{scenarios},
on a network of three detectors with aLIGO sensitivity. The network SNR
threshold of $12$ yields rates which are roughly comparable with rates
computed using an SNR threshold of $8$ in a single aLIGO detector as proxy for the network.
The actual detection thresholds are a complicated function
of network configuration, the level and frequency of non-Gaussian,
non-stationary excursions in the noise, and search pipeline
sensitivity to different source types. Therefore, our simple
thresholds are only meant to yield rough estimates of detection rates,
and the focus should be on relative rates for different source types
and model assumptions rather than absolute numbers. Finally, we note
that the sensitivity of advanced detectors will gradually improve
during commissioning, and several years will pass before they reach
the sensitivity we have assumed above \citep[for an approximate time
line, see][]{scenarios}.
The detection rates computed by assuming an SNR threshold of $8$ in a single aLIGO
detector as proxy for the network allow for a direct comparison with the rate ranges
compiled in \citep{2010CQGra..27q3001A}, which used the same detectability criterion.
\citet{2010CQGra..27q3001A} incorporated a number of population synthesis studies and
Galactic binary pulsar observations, but did not include some of the factors considered
in the present study, such as cosmology and variations in metallicity distributions and star
formation rates with redshift. We find that our predicted detection rates for NS-NS and
BH-BH binaries fall within the ranges given in \citep{2010CQGra..27q3001A} for all models
and both metallicity distribution choices considered in the present work. For BH-NS binaries,
the same holds for all models and metallicity choices except for the high BH kick model, which
yields BH-NS detection rates below the range quoted in \citep{2010CQGra..27q3001A}.
We note that uncertainties in waveform systematics and detection
criteria pale in comparison to uncertainties in stellar and binary
evolution. We consider the most important uncertainties to be the
progress and outcome of the CE phase, the SN explosion mechanism and
the magnitude of BH natal kicks. The four binary evolution models
discussed in this study explore these uncertainties, resulting in a
wide range of mass distributions and event rates. Changing other
parameters such as the initial binary mass distribution or varying the
mass escaping the systems during mass transfer episodes would also
influence the resulting distributions and rates
\citep{2005ApJ...620..385O,2008ApJ...675..566O,roskb}.
The properties of the DCO populations produced in our various models
are sufficiently differentiated that it may be possible to constrain
or rule out some of the input physics based on observed populations.
For example, a lack of significant number of detections will disfavor the Optimistic CE model, in
which we allow for CE events with HG donors and thus find very high detection rates.
This will indicate how (if at all) CE develops for HG stars.
If BH-BH systems are not detected far more
frequently than other DCO types, a likely explanation is that BHs
receive significant natal kicks disrupting their binaries. A detailed
comparison of detection rates with current LIGO upper limits can be
found in~\citet{comparison}. As detections accumulate, a well measured
chirp mass distribution could allow us to distinguish between the
Rapid and Delayed SN engine models, which generate continuous and
gapped chirp mass distribution of DCOs, respectively. The number of
detections needed to distinguish between the Rapid and Delayed SN
engines will be discussed in future work (Dominik et al. 2014, in
preparation).
\acknowledgements
We thank a number of LIGO and Virgo collaboration colleagues,
particularly Thomas Dent, David Shoemaker, Stephen Fairhurst and Peter
Saulson,
for advice on the manuscript. We thank the N. Copernicus Astronomical
Centre in Warsaw, Poland, and the University of Texas at Brownsville,
for providing computational resources. The authors acknowledge the
Texas Advanced Computing Center (TACC) at The University of Texas at
Austin for providing computational resources.
KB acknowledges support from a Polish Science Foundation “Master2013”
Subsidy, Polish NCN grant SONATA BIS 2, NASA Grant Number NNX09AV06A and NSF
Grant Number HRD 1242090 awarded to the Center for Gravitational Wave
Astronomy at U.T. Brownsville.
MD acknowledges support from the National Science Center grant
DEC-2011/01/N/ST9/00383.
EB acknowledges support
from National Science Foundation CAREER Grant PHY-1055103.
ROS was supported by NSF award PHY-0970074 and the UWM Research Growth
Initiative.
DEH acknowledges support from National Science Foundation CAREER grant
PHY-1151836. He was also supported in part by the Kavli Institute for
Cosmological Physics at the University of Chicago through NSF grant
PHY-1125897 and an endowment from the Kavli Foundation and its founder
Fred Kavli.
TB was supported by the DPN/N176/VIRGO/2009 grant and the
DEC-2013/01/ASPERA/ST9/00001 from the National Science Center, Poland.
FP was supported by STFC Grant No.~ST/L000342/1.
This work was supported in part by the National Science Foundation under
Grant No. PHYS-1066293 and the hospitality of the Aspen Center for Physics (KB).
The study was also sponsored by the National Science Center grant Sonata Bis 2
(DEC-2012/07/E/ST9/01360).
\bibliographystyle{aa}
|
\section{Introduction}
Data science has become a driving force for many scientific studies. As datasets get bigger and the methods of analysis become more complex, the need for reproducibility has increased significantly \citep{SLP14}. A minimal requirement of reproducibility is that one can reach similar results based on independently generated datasets. The issue of reproducibility has drawn much attention in the scientific community (see a special issue of \textit{Nature}\footnote{at http://www.nature.com/nature/focus/reproducibility/}); Marcia McNutt, the Editor-in-Chief of \textit{Science}, pointed out that ``reproducing an experiment is one important approach that scientists use to gain confidence in their conclusions.'' In other words, if conclusions cannot be reproduced, the credit of the researchers, along with the scientific conclusions themselves, will be in jeopardy.
\subsection{Stability}
Statistics as a subject can help improve reproducibility in many ways. One particular aspect we stress in this article is the stability of a statistical procedure used in the analysis. According to \cite{Y13}, ``reproducibility manifests itself in stability of statistical results relative to `reasonable' perturbations to data and to the model used.'' An instable statistical method leads to the possibility that a correct scientific conclusion is not reproducible, and hence is not recognized, or even falsely discredited.
Stability has indeed received much attention in statistics. However, few work has focused on stability itself. Many works instead view stability as a tool for other purposes. For example, in clustering problems, \citet{BEG02} introduced the clustering instability to assess the quality of a clustering algorithm; \citet{W10} used the clustering instability as a criterion to select the number of clusters. In high-dimensional regression, \citet{MB10} proposed stability selection procedures for variable selection; \citet{LRW10} and \citet{SWF13} applied stability for tuning parameter selection. For more applications, see the use of stability in model selection \citep{B96}, analyzing the effect of bagging \citep{BY02}, and deriving the generalization error bound \citep{BE02,EEP05}. While successes of stability have been reported in the aforementioned works, to the best of our knowledge, there has been little systematic methodological and theoretical study of stability itself in the classification context.
On the other hand, we are aware that ``a study can be reproducible but still be wrong''\footnote{http://simplystatistics.org/2014/06/06/the-real-reason-reproducible-research-is-important/}. So can a classification method be stable but inaccurate. Thus, in this article, stability is not meant to replace classification accuracy, which is the primary goal for much of the research work on classification. However, an irreproducible or instable study will definitely reduce its chance of being accepted by the scientific community, no matter how accurate it is. Hence, it is ideal for a method to be both accurate and stable, a goal of the current article.
Moreover, in certain practical domains of classification, stability can be as important as accuracy. This is because providing a stable prediction plays a crucial role on users' trust on a system. For example, Internet streaming service provider Netflix has a movie recommendation system based on complex supervised learning algorithms. In this application, if two consecutively recommended movies are from two totally different genres, the viewers can immediately perceive such instability, and have a bad user experience with the service \citep{AZ10}.
\subsection{Overview}
The $k$-nearest neighbor ($k$NN) classifier \citep{FH51,CH67} is one of the most popular nonparametric classification methods, due to its conceptual simplicity and powerful prediction capability. In the literature, extensive research have been done to justify various nearest neighbor classifiers based on the risk, which measures the inaccuracy of a classifier \citep{DW77, S77, G81, DGKL94, SV98, BCG10}. We refer the readers to \citet{DGL96} for a comprehensive study. Recently, \citet{S12} has proposed an optimal weighted nearest neighbor (OWNN) classifier. Like most other existing nearest neighbor classifiers, OWNN focuses on the risk without paying attention to the classification stability.
\begin{figure}[t!]
\begin{center}
\includegraphics[scale=0.6]{stabknn_regret_cis_new}\vspace{-1em}
\caption{\label{stabknn_regret_cis} \footnotesize Regret and CIS of the $k$NN classifier. From top to bottom, each circle represents the $k$NN classifier with $k\in \{1,2,\dots,20\}$. The red square corresponds to the classifier with the minimal regret and the classifier depicted by the blue triangle improves it to have a lower CIS.}
\end{center}
\end{figure}
In this article, we define a general measure of stability for a classification method, named as \textit{Classification Instability} (CIS). It characterizes the sampling variability of the prediction. An important result we show is that the asymptotic CIS of any weighted nearest neighbor classifier (a generalization of $k$NN), denoted as WNN, turns out to be proportional to the Euclidean norm of its weight vector. This rather concise form is crucial in our methodological development and theoretical analysis. To illustrate the relation between risk and CIS, we apply the $k$NN classifier to a toy example (see details in Section \ref{sec:validation}) and plot in Figure \ref{stabknn_regret_cis} the regret (that is, the risk minus a constant known as the Bayes risk) versus CIS, calculated according to Proposition \ref{thm:regret} and Theorem \ref{thm:CIS} in Section~\ref{sec:snn0}, for different $k$. As $k$ increases, the classifier becomes more and more stable, while the regret first decreases and then increases. In view of the $k$NN classifier with the minimal regret, marked as the red square in Figure \ref{stabknn_regret_cis}, one may have the impression that there are other $k$ values with similar regret but much smaller CIS, such as the one marked as the blue triangle shown in the plot.
Inspired by Figure \ref{stabknn_regret_cis}, we propose a novel method called stabilized nearest neighbor (SNN) classifier, which takes the stability into consideration. The SNN procedure is constructed by minimizing the CIS of WNN over an acceptable region where the regret is small, indexed by a tuning parameter. SNN encompasses the OWNN classifier as a special case.
To understand the theoretical property of SNN, we establish a sharp convergence rate of CIS for general plug-in classifiers. This sharp rate is slower than but approaching $n^{-1}$, shown by adapting the framework of \citet{AT07}. Furthermore, the proposed SNN method is shown to achieve both the minimax optimal rate in the regret established in the literature, and the sharp rate in CIS established in this article.
\begin{figure}[t!]
\begin{center}
\vspace{-1em}\includegraphics[scale=0.6]{regret_cis_asymptotic_new}\vspace{-1em}
\caption{\label{regret_cis_comparison} \footnotesize Regret and CIS of $k$NN, OWNN, and SNN procedures for a bivariate normal example. The top three lines represent CIS's of $k$NN, OWNN, and SNN. The bottom three lines represent regrets of $k$NN, SNN, and OWNN. The sample size shown on the x-axis is in the $\log_{10}$ scale.}
\end{center}
\end{figure}
To further illustrate the advantages of the SNN classifier, we offer a comprehensive asymptotic comparison among various classifiers, through which new insights are obtained. It is theoretically verified that the CIS of our SNN procedure is much smaller than those of others. Figure \ref{regret_cis_comparison} shows the regret and CIS of $k$NN, OWNN, and SNN for a bivariate example (see details in Section \ref{sec:validation}). Although OWNN is \textit{theoretically} the best in regret, its regret curve appear to overlap with that of SNN. On the other hand, the SNN procedure has a noticeably smaller CIS than OWNN. A compelling message is that with almost the same accuracy, our SNN could greatly improve stability. In the finite sample case, extensive experiments confirm that SNN has a significant improvement in CIS, and sometimes even improves accuracy slightly. Such appealing results are supported by our theoretical finding (in Corollary \ref{cor:snn_ownn}) that the regret of SNN approaches that of OWNN at a faster rate than the rate at which the CIS of OWNN approaches that of SNN, where both rates are shown to be sharp. As a by-product, we also show that OWNN is more stable than $k$NN and bagged nearest neighbor (BNN) classifiers.
The rest of the article is organized as follows. Section~\ref{sec:cis} defines CIS for a general classification method. In Section~\ref{sec:snn0}, we study the stability of the nearest neighbor classifier, and propose a novel SNN classifier. The SNN classifier is shown to achieve an established sharp rate in CIS and the minimax optimal rate in regret in Section~\ref{sec:thms}. Section~\ref{theoretical_compare} presents a thorough theoretical comparison of regret and CIS between the SNN classifier and other nearest neighbor classifiers. Section \ref{sec:algorithm} discusses the issue of tuning parameter selection, followed by numerical studies in Section~\ref{sec:exp}. We conclude the article in Section~\ref{discussion}. The appendix and supplementary materials are devoted to technical proofs.
\section{Classification Instability}\label{sec:cis}
Let $(X,Y)\in{\mathbb R}^d \times\{1,2\}$ be a random couple with a joint distribution $P$. We regard $X$ as a $d$-dimensional vector of features for an object and $Y$ as a label indicating that the object belongs to one of two classes. Denote the prior class probability as $\pi_1={\mathbb P}(Y=1)$, where $\mathbb P$ is the probability with respect to $P$, and the distribution of $X$ given $Y=r$ as $P_r$ with $r=1,2$. The marginal distribution of $X$ can be written as $\bar{P}=\pi_1 P_1 + (1-\pi_1)P_2$. For a classifier $\phi: {\mathbb R}^d \mapsto\{1,2\}$, the risk of $\phi$ is defined as $R(\phi) = \mathbb P(\phi(X)\ne Y)$. It is well known that the Bayes rule, denoted as $\phi^{\textrm{Bayes}}$, minimizes the above risk. Specifically, $\phi^{\textrm{Bayes}}(x)=1+\ind{\eta(x)< 1/2}$, where $\eta(x)={\mathbb P}(Y=1|X=x)$ and $\ind{\cdot}$ is the indicator function. In practice, a classification procedure $\Psi$ is applied to a training data set ${\cal D}= \{(X_i,Y_i), i=1,\ldots,n\}$ to produce a classifier $\widehat\phi_{n}=\Psi({\cal D})$. We define the risk of the procedure $\Psi$ as $\mathbb E_{{\cal D}}[R(\widehat\phi_{n})]$, and the regret of $\Psi$ as $\mathbb E_{{\cal D}}[R(\widehat\phi_{n})]-R(\phi^{\textrm{Bayes}})$, where $\mathbb E_{{\cal D}}$ denotes the expectation with respect to the distribution of $\cal D$, and $R(\phi^{\textrm{Bayes}})$ is called Bayes risk. Both the risk and regret describe the inaccuracy of a classification method. In practice, for a classifier $\phi$, the classification error for a test data can be calculated as an empirical version of $R(\phi)$.
For a classification procedure, it is desired that, with high probability, classifiers trained from different samples yield the same prediction for the same object. Our first step in formalizing the classification instability is to define the distance between two generic classifiers $\phi_1$ and $\phi_2$, which measures the level of disagreement between them.
\begin{defi}
(Distance between Classifiers) Define the distance between two classifiers $\phi_1$ and $\phi_2$ as $d(\phi_1,\phi_2) = {\mathbb P}(\phi_1(X) \ne \phi_2(X))$.
\end{defi}
We next define the classification instability (CIS). Throughout the article, we denote ${\cal D}_1$ and ${\cal D}_2$ as two i.i.d. copies of the training sample ${\cal D}$. For ease of notation, we have suppressed the dependence of $\textrm{CIS}(\Psi)$ on the sample size $n$ of $\mathcal D$.
\begin{defi}(Classification Instability)
Define the classification instability of a classification procedure $\Psi$ as
\begin{equation}
\textrm{CIS}(\Psi) = {\mathbb E}_{{\cal D}_1,{\cal D}_2}\Big[ d(\widehat\phi_{n1},\widehat\phi_{n2})\Big]\label{CIS}
\end{equation}
where $\widehat\phi_{n1}=\Psi(\mathcal D_1)$ and $\widehat\phi_{n2}=\Psi(\mathcal D_2)$ are the classifiers obtained by applying the classification procedure $\Psi$ to samples ${\cal D}_1$ and ${\cal D}_2$.
\end{defi}
Intuitively, CIS is an average probability that the same object is classified to two different classes in two separate runs of a learning algorithm. By definition, $0\le \textrm{CIS}(\Psi) \le 1$, and a small $\textrm{CIS}(\Psi)$ represents a stable classification procedure $\Psi$.
\section{Stabilized Nearest Neighbor Classifier}\label{sec:snn0}
\subsection{Review of WNN}\label{sec:WNN}
For any fixed $x$, let $(X_{(1)},Y_{(1)}),\ldots, (X_{(n)},Y_{(n)})$ be a sequence of observations with ascending distance to $x$. For a nonnegative weight vector $\boldsymbol{w}_n=(w_{ni})_{i=1}^n$ satisfying $\sum_{i=1}^n w_{ni} = 1$, a WNN classifier $\widehat{\phi}_{n}^{\boldsymbol{w}_n}$ predicts the label of $x$ as $\widehat{\phi}_{n}^{\boldsymbol{w}_n}(x)=1+\ind{\sum_{i=1}^n w_{ni}\ind{Y_{(i)}=1}< 1/2}$. \citet{S12} revealed a nice asymptotic expansion formula for the regret of WNN.
\begin{pro}
\label{thm:regret}
\citep{S12} Under Assumptions (A1)--(A4) defined in Appendix \ref{sec:assumptions}, for each $\beta\in (0,1/2)$, we have, as $n\rightarrow \infty$,
\begin{equation}
\textrm{Regret}(\textrm{WNN}) = \left\{B_1 \sum_{i=1}^n w_{ni}^2 + B_2 \Big (\sum_{i=1}^n \frac{\alpha_i w_{ni}}{n^{2/d}}\Big)^2 \right\}\{1+o(1)\},\label{regret}
\end{equation}
uniformly for $\boldsymbol{w}_n\in W_{n,\beta}$ with $W_{n,\beta}$ defined in Appendix \ref{sec:defwnb}, where $\alpha_i=i^{1+\frac{2}{d}}-(i-1)^{1+\frac{2}{d}}$, and constants $B_1$ and $B_2$ are defined in Appendix \ref{sec:defwnb}.
\end{pro}
\citet{S12} further derived a weight vector that minimizes the asymptotic regret $(\ref{regret})$ which led to the optimal weighted nearest neighbor (OWNN) classifier.
\subsection{Asymptotically Equivalent Formulation of CIS}\label{sec:asy_CIS}
Denote two resulting WNN classifiers trained on ${\cal D}_1$ and ${\cal D}_2$ as $\widehat{\phi}_{n1}^{\boldsymbol{w}_n}(x)$ and $\widehat{\phi}_{n2}^{\boldsymbol{w}_n}(x)$ respectively. With a slight abuse of notation, we denote the CIS of a WNN classification procedure by $\textrm{CIS}({\makebox{WNN}})$. According to the definition in (\ref{CIS}), classification instability of a WNN procedure is
$\textrm{CIS}(\textrm{WNN})= {\mathbb P}_{{\cal D}_1,{\cal D}_2,X}\Big(\widehat{\phi}_{n1}^{\boldsymbol{w}_n}(X) \ne \widehat{\phi}_{n2}^{\boldsymbol{w}_n}(X)\Big).$
Theorem \ref{thm:CIS} provides an asymptotic expansion formula for the CIS of WNN in terms of its weight vector $\boldsymbol{w}_n$.
\begin{theorem}
\label{thm:CIS}
(Asymptotic CIS) Under Assumptions (A1)--(A4) defined in Appendix \ref{sec:assumptions}, for each $\beta\in (0,1/2)$, we have, as $n\rightarrow \infty$,
\begin{equation}
\textrm{CIS}(\textrm{WNN}) = B_3 \Big(\sum_{i=1}^n w_{ni}^2\Big)^{1/2} \{1+o(1)\}, \label{eq:asy_CIS}
\end{equation}
uniformly for all $\boldsymbol{w}_n\in W_{n,\beta}$ with $W_{n,\beta}$ defined in Appendix \ref{sec:defwnb}, where the constant $B_3 = 4B_1/\sqrt{\pi}>0$ with $B_1$ defined in Appendix \ref{sec:defwnb}.
\end{theorem}
Theorem \ref{thm:CIS} demonstrates that the asymptotic CIS of a WNN procedure is proportional to $(\sum_{i=1}^n w_{ni}^2)^{1/2}$. For example, for the $k$NN procedure (that is the WNN procedure with $w_{ni}=k^{-1}\ind{1\le i\le k}$), its CIS is asymptotically $B_3\sqrt{1/k}$. Therefore, a larger value of $k$ leads to a more stable $k$NN procedure, which was seen in Figure \ref{stabknn_regret_cis}. Furthermore, we note that the CIS expansion in (\ref{eq:asy_CIS}) is related to the first term in (\ref{regret}). The expansions in (\ref{regret}) and (\ref{eq:asy_CIS}) allow precise calibration of regret and CIS. This delicate connection is important in the development of our SNN procedure.
\subsection{Stabilized Nearest Neighbor Classifier}
\label{sec:SNN}
To stabilize WNN, we consider a weight vector which minimizes the CIS over an acceptable region where the classification regret is less than some constant $c_1>0$, that is,
\begin{align}
\min_{\boldsymbol{w}_n}~~ &\textrm{CIS}(\textrm{WNN}), \label{original}\\
\textrm{subject to}~~ &\textrm{Regret}(\textrm{WNN}) \le c_1,~\sum_{i=1}^n w_{ni} =1,~\boldsymbol{w}_{n} \ge 0. \nonumber
\end{align}
By a non-decreasing transformation, we change the objective function in (\ref{original}) to $\textrm{CIS}^2(\textrm{WNN})$. Furthermore, considering the Lagrangian formulation, we can see that (\ref{original}) is equivalent to minimizing $\textrm{Regret}(\textrm{WNN}) + \lambda_0\textrm{CIS}^2(\textrm{WNN})$ subject to the constraints that $\sum_{i=1}^n w_{ni} =1$ and $\boldsymbol{w}_{n} \ge 0$, where $\lambda_0>0$. The equivalence is ensured by the expansions (\ref{regret}) and (\ref{eq:asy_CIS}) in Proposition \ref{thm:regret} and Theorem \ref{thm:CIS}, and the fact that both the objective function and the constraints are convex in the variable vector $\boldsymbol{w}_n$. The resulting optimization is
\begin{align}
\min_{\boldsymbol{w}_n}~~ &\left(\sum_{i=1}^n \frac{\alpha_i w_{ni}}{n^{2/d}}\right)^2 + \lambda \sum_{i=1}^n w_{ni}^2,\label{formal}\\
\textrm{subject to}~~ &\sum_{i=1}^n w_{ni} =1,~\boldsymbol{w}_{n} \ge 0, \nonumber
\end{align}
where $\lambda = (B_1+\lambda_0B_3^2)/B_2$ depends on constants $B_1$ and $B_2$ and $\lambda_0$. When $\lambda\rightarrow \infty$, (\ref{formal}) leads to the most stable but trivial $k$NN classifier with $k=n$. The classifier in (\ref{formal}) with $\lambda\downarrow B_1/B_2$ (\textit{i.e.}, $\lambda_0\downarrow 0$) approaches the OWNN classifier considered in \citet{S12}. Note that the two terms $(n^{-2/d}\sum_{i=1}^n \alpha_i w_{ni})^2$ and $\sum_{i=1}^n w_{ni}^2$ in (\ref{formal}) represent the bias and variance terms of the regret expansion given in Proposition \ref{thm:regret} \citep{S12}. By varying the weights of these two terms through $\lambda$, we are able to stabilize a nearest neighbor classifier. Moreover, the stabilized classifier achieves desirable convergence rates in both regret and CIS; see Section~\ref{sec:thms}.
Theorem \ref{thm:optimal} gives the optimal weight $w_{ni}^*$ with respect to the optimization (\ref{formal}). We formally define the stabilized nearest neighbor (SNN) classifier as the WNN classifier with the optimal weight $w_{ni}^*$.
\begin{theorem}
\label{thm:optimal}
(Optimal Weight) For any fixed $\lambda>0$, the minimizer of $(\ref{formal})$ is
\[ w_{ni}^* = \left\{ \begin{array}{ll}
\frac{1}{k^*}\left(1+\frac{d}{2}-\frac{d}{2(k^*)^{2/d}}\alpha_i\right), & \mbox{for $i=1,\ldots,k^*$},\\
0, & \mbox{for $i = k^*+1, \ldots, n$},\end{array} \right. \]
where $\alpha_i=i^{1+\frac{2}{d}}-(i-1)^{1+\frac{2}{d}}$ and $k^*= \lfloor \{\frac{d(d+4)}{2(d+2)}\}^{\frac{d}{d+4}}\lambda^{\frac{d}{d+4}}n^{\frac{4}{d+4}}\rfloor.$
\end{theorem}
The SNN classifier encompasses the OWNN classifier as a special case when $\lambda=B_1/B_2$.
The computational complexity of our SNN classifier is comparable to that of existing nearest neighbor classifiers. If we preselect a value for $\lambda$, SNN requires no training at all. The testing time consists of two parts: an $O(n)$ complexity for the computation of $n$ distances, where $n$ is the size of training data; and an $O(n\log n)$ complexity for sorting $n$ distances. The $k$NN classifier, for example, shares the same computational complexity. In practice, $\lambda$ is not predetermined and we may treat it as a tuning parameter, whose optimal value is selected via cross validation. See Algorithm 1 in Section~\ref{sec:algorithm} for details. We will show in Section~\ref{sec:algorithm} that the complexity of tuning in SNN is also comparable to existing methods.
\section{Theoretical Properties}\label{sec:thms}
\subsection{A Sharp Rate of CIS}\label{sec:sharprate}
Motivated by \citet{AT07}, we establish a sharp convergence rate of CIS for a general plug-in classifier. A plug-in classification procedure $\Psi$ first estimates the regression function $\eta(x)$ by $\widehat{\eta}_n(x)$, and then plugs it into the Bayes rule, that is, $\widehat\phi_n(x) = 1+ \ind{\widehat{\eta}_n(x)< 1/2}$.
The following \textit{margin condition} \citep{T04} is assumed for deriving the upper bound of the convergence rate, while two additional conditions are required for showing the lower bound. A distribution function $P$ satisfies the \textit{margin condition} if there exist constants $C_0>0$ and $\alpha\ge 0$ such that for any $\epsilon>0$,
\begin{align}
\mathbb P_X(0<|\eta(X)-1/2|\le \epsilon)\le C_0 \epsilon^{\alpha}.\label{margin}
\end{align}
The parameter $\alpha$ characterizes the behavior of the regression function $\eta$ near $1/2$, and a larger $\alpha$ implies a lower noise level and hence an easier classification scenario.
The second condition is on the smoothness of $\eta(x)$. Specifically, we assume that $\eta$ belongs to a \textit{H{\"o}lder class of functions} $\Sigma(\gamma,L,\mathbb R^d)$ (for some fixed $L, \gamma>0$) containing the functions $g:\mathbb R^d\rightarrow \mathbb R$ that are $\lfloor \gamma\rfloor$ times continuously differentiable and satisfy, for any $x,x'\in \mathbb R^d$,
$|g(x') - g_x(x')| \le L\|x-x'\|^{\gamma},$
where $\lfloor \gamma \rfloor$ is the largest integer not greater than $\gamma$, $g_x$ is the Taylor polynomial series of degree $\lfloor \gamma\rfloor$ at $x$, and $\|\cdot\|$ is the Euclidean norm.
Our last condition assumes that the marginal distribution $\bar P$ satisfies the \textit{strong density assumption}, defined in Supplementary \ref{sec:strongdensity}.
We first derive the rate of convergence of CIS by assuming an exponential convergence rate of the corresponding regression function estimator.
\begin{theorem}
\label{thm:upperCIS}
(Upper Bound) Let $\widehat{\eta}_n$ be an estimator of the regression function $\eta$ and let ${\cal R}\subset \mathbb R^d$ be a compact set. Let $\mathcal P$ be a set of probability distributions supported on ${\cal R} \times \{1,2\}$ such that for some constants $C_1, C_2>0$, some positive sequence $a_n\rightarrow \infty$, and almost all $x$ with respect to $\bar{P}$,
\begin{align}
\sup_{P\in {\cal P}} \mathbb P_{\cal D}\Big(|\widehat{\eta}_n(x)-\eta(x)| \ge \delta\Big) \le C_1 \exp(-C_2 a_n \delta^2) \label{exponential}
\end{align}
holds for any $n> 1$ and $\delta>0$, where $\mathbb P_{\cal D}$ is the probability with respect to $P^{\otimes n}$. Furthermore, if all the distributions $P\in\cal P$ satisfy the margin condition for a constant $C_0$, then the plug-in classification procedure $\Psi$ corresponding to $\widehat{\eta}_n$ satisfies
$$
\sup_{P\in {\cal P}} \textrm{CIS}(\Psi) \le C a_n^{-\alpha/2},
$$
for any $n> 1$ and some constant $C>0$ depending only on $\alpha,C_0, C_1$, and $C_2$.
\end{theorem}
It is worth noting that the condition in $(\ref{exponential})$ holds for various types of estimators. For example, Theorem 3.2 in \citet{AT07} showed that the local polynomial estimator satisfies $(\ref{exponential})$ with $a_n=n^{2\gamma/(2\gamma+d)}$ when the bandwidth is of the order $n^{-1/(2\gamma+d)}$. In addition, Theorem \ref{thm:upperCISsnn} in Section \ref{sec:ratesnn} implies that $(\ref{exponential})$ holds for the newly proposed SNN classifier with the same $a_n$. Hence, in both cases, the upper bound is of the order $n^{-\alpha\gamma/(2\gamma+d)}$.
We next derive the lower bound of CIS in Theorem \ref{thm:lowerCIS}. As will be seen, this lower bound implies that the obtained rate of CIS, that is, $n^{-\alpha\gamma/(2\gamma+d)}$, cannot be further improved for the plug-in classification procedure.
\begin{theorem}
\label{thm:lowerCIS}
(Lower Bound) Let ${\cal P}_{\alpha,\gamma}$ be a set of probability distributions supported on ${\cal R} \times \{1,2\}$ such that for any $P\in{\cal P}_{\alpha,\gamma}$, $P$ satisfies the margin condition $(\ref{margin})$, the regression function $\eta(x)$ belongs to the H\"older class $\Sigma(\gamma,L,\mathbb R^d)$, and the marginal distribution $\bar{P}$ satisfies the strong density assumption. Suppose further that ${\cal P}_{\alpha,\gamma}$ satisfies $(\ref{exponential})$ with $a_n=n^{2\gamma/(2\gamma+d)}$ and $\alpha\gamma\le d$. We have
$$
\sup_{P\in {\cal P}_{\alpha,\gamma}} \textrm{CIS}(\Psi) \ge C' n^{-\alpha\gamma/(2\gamma+d)},
$$
for any $n> 1$ and some constant $C'>0$ independent of $n$.
\end{theorem}
Theorems \ref{thm:upperCIS} and \ref{thm:lowerCIS} together establish a sharp convergence rate of the CIS for the general plug-in classification procedure on the set ${\cal P}_{\alpha,\gamma}$. The requirement $\alpha\gamma\le d$ in Theorem \ref{thm:lowerCIS} implies that $\alpha$ and $\gamma$ cannot be large simultaneously. As pointed out in \citet{AT07}, this is intuitively true because a very large $\gamma$ implies a very smooth regression function $\eta$, while a large $\alpha$ implies that $\eta$ cannot stay very long near $1/2$, and hence when $\eta$ hits $1/2$, it should take off quickly. Lastly, we note that this rate is slower than $n^{-1}$, but approaches $n^{-1}$ as the dimension $d$ increases when $\alpha\gamma=d$.
As a reminder, \citet{AT07} established the minimax optimal rate of regret as $n^{-(\alpha+1)\gamma/(2\gamma+d)}$.
\subsection{Optimal Convergence Rates of SNN}
\label{sec:ratesnn}
In this subsection, we demonstrate that SNN attains the established sharp convergence rate in CIS in the previous subsection, as well as the minimax optimal convergence rate in regret. We further show the asymptotic difference between SNN and OWNN.
In Theorem \ref{thm:upperCISsnn} and Corollary \ref{cor:snn_ownn} below, we consider $\textrm{SNN}$ with $k^* \asymp n^{2\gamma/(2\gamma+d)}$ in Theorem \ref{thm:optimal}, where $a_n \asymp b_n$ means the ratio sequence $a_n/b_n$ stays away from zero and infinity as $n\rightarrow\infty$. Note that under Assumptions (A1)--(A4) defined in Appendix \ref{sec:assumptions}, we have $\gamma=2$ and hence $k^* \asymp n^{4/(4+d)}$, which agrees with the formulation in Theorem \ref{thm:optimal}.
\begin{theorem}
\label{thm:upperCISsnn}
For any $\alpha\ge 0$ and $\gamma\in (0,2]$, the SNN procedure with any fixed $\lambda>0$ satisfies
\begin{eqnarray*}
\sup_{P\in {\cal P}_{\alpha,\gamma}} \textrm{Regret}(\textrm{SNN}) &\le& \tilde{C} n^{-(\alpha+1)\gamma/(2\gamma+d)},\\
\sup_{P\in {\cal P}_{\alpha,\gamma}} \textrm{CIS}(\textrm{SNN}) &\le& C n^{-\alpha\gamma/(2\gamma+d)},
\end{eqnarray*}
for any $n> 1$ and some constants $\tilde{C}, C>0$, where ${\cal P}_{\alpha,\gamma}$ is defined in Theorem~\ref{thm:lowerCIS}.
\end{theorem}
Corollary \ref{cor:snn_ownn} below further investigates the difference between the SNN procedure (with $\lambda\ne B_1/B_2$) and the OWNN procedure in terms of both regret and CIS.
\begin{corollary}
\label{cor:snn_ownn}
For any $\alpha\ge 0$, $\gamma\in (0,2]$, we have, when $\lambda\ne B_1/B_2$,
\begin{eqnarray}
\sup_{P\in {\cal P}_{\alpha,\gamma}} \Big\{\textrm{Regret}(\textrm{SNN})-\textrm{Regret}(\textrm{OWNN})\Big\} &\asymp& n^{-(1+\alpha)\gamma/(2\gamma+d)},\nonumber\\
\sup_{P\in {\cal P}_{\alpha,\gamma}} \Big\{ \textrm{CIS}(\textrm{OWNN}) - \textrm{CIS}(\textrm{SNN})\Big\} &\asymp& n^{-\alpha\gamma/(2\gamma+d)}, \label{eqn:cisdifference}
\end{eqnarray}
where ${\cal P}_{\alpha,\gamma}$ is defined in Theorem~\ref{thm:lowerCIS}.
\end{corollary}
Corollary \ref{cor:snn_ownn} implies that the regret of SNN approaches that of the OWNN (from above) at a faster rate than the CIS of OWNN approaches that of the SNN procedure (from above). This means that SNN can have a significant improvement in CIS over the OWNN procedure while obtaining a comparable classification accuracy. This observation will be supported by the experimental results in Section \ref{sec:simulation}.
\begin{remark}
Under Assumptions (A1)--(A4) in Section \ref{sec:assumptions}, which implicitly implies that $\alpha=1$, and the assumption that $\gamma=2$, the conclusion in $(\ref{eqn:cisdifference})$ can be strengthened to that for any $P\in {\cal P}_{1,2}$, $\textrm{CIS}(\textrm{OWNN}) - \textrm{CIS}(\textrm{SNN}) \asymp n^{-2/(d+4)}$. It indicates that SNN's improvement in CIS is at least $n^{-2/(d+4)}$ in this scenario.
\end{remark}
\section{Asymptotic Comparisons}\label{theoretical_compare}
In this section, we first conduct an asymptotic comparison of CIS among existing nearest neighbor classifiers, and then demonstrate that SNN significantly improves OWNN in CIS.
\subsection{CIS Comparison of Existing Methods}\label{sec:ciscom}
We compare $k$NN, OWNN and the bagged nearest neighbor (BNN) classifier. The $k$NN classifier is a special case of the WNN classifier with weight $w_{ni}=1/k$ for $i=1,\ldots,k$ and $w_{ni}=0$ otherwise. Another special case of the WNN classifier is the BNN classifier. After generating subsamples from the original data set, the BNN classifier applies 1-nearest neighbor classifier to each bootstrapped subsample and returns the final prediction by majority voting. If the resample size $m$ is sufficiently smaller than $n$, \textit{i.e.}, $m\rightarrow \infty$ and $m/n\rightarrow 0$, the BNN classifier is shown to be a consistent classifier \citep{HS05}. In particular, \citet{HS05} showed that, for large $n$, the BNN classifier (with or without replacement) is approximately equivalent to a WNN classifier with the weight $w_{ni} = q(1-q)^{i-1}/[1-(1-q)^n]$ for $i=1,\ldots,n$, where $q$ is the resampling ratio $m/n$.
We denote the CIS of the above classification procedures as $\textrm{CIS}(\textrm{$k$NN})$, $\textrm{CIS}(\textrm{BNN})$ and $\textrm{CIS}(\textrm{OWNN})$. Here $k$ in the $k$NN classifier is selected as the one minimizing the regret \citep{HPS08}. The optimal $q$ in the BNN classifier and the optimal weight in the OWNN classifier are both calculated based on their asymptotic relations with the optimal $k$ in $k$NN, which were defined in (2.9) and (3.5) of \citet{S12}. Corollary \ref{cor:cisratio} gives the pairwise CIS ratios of these classifiers. Note that these ratios depend on the feature dimension $d$ only.
\begin{corollary}
\label{cor:cisratio}
Under Assumptions (A1)-(A4) defined in Appendix \ref{sec:assumptions} and the assumption that $B_2$ defined in Appendix \ref{sec:defwnb} is positive, we have, as $n\rightarrow \infty$,
\begin{eqnarray*}
\frac{\textrm{CIS}(\textrm{OWNN})}{\textrm{CIS}(\textrm{$k$NN})} &\longrightarrow& 2^{2/(d+4)}\Big(\frac{d+2}{d+4}\Big)^{(d+2)/(d+4)},\\
\frac{\textrm{CIS}(\textrm{BNN})}{\textrm{CIS}(\textrm{$k$NN})} &\longrightarrow& 2^{-2/(d+4)}\Gamma(2+2/d)^{d/(d+4)}, \\
\frac{\textrm{CIS}(\textrm{BNN})}{\textrm{CIS}(\textrm{OWNN})} &\longrightarrow& 2^{-4/(d+4)}\Gamma(2+2/d)^{d/(d+4)}\Big(\frac{d+4}{d+2}\Big)^{(d+2)/(d+4)}.
\end{eqnarray*}
\end{corollary}
\begin{figure}[!b]
\begin{center}\vspace{-3em}
\includegraphics[scale=0.6]{cisratio_ownn_knn_bnn_new}\vspace{-1em}
\caption{\label{CISratio_ownn_knn_bnn} \footnotesize Pairwise CIS ratios between $k$NN, BNN and OWNN for different feature dimension $d$.}
\end{center}
\end{figure}
The limiting CIS ratios in Corollary \ref{cor:cisratio} are plotted in Figure \ref{CISratio_ownn_knn_bnn}. A major message herein is that the OWNN procedure is more stable than the $k$NN and BNN procedures for any $d$. The largest improvement of the OWNN procedure over $k$NN is achieved when $d=4$ and the improvement diminishes as $d\rightarrow \infty$. The CIS ratio of BNN over $k$NN equals $1$ when $d=2$ and is less than $1$ when $d>2$, which is consistent with the common perception that bagging can generally reduce the variability of the nearest neighbor classifiers. Similar phenomenon has been shown in the ratio of their regrets \citep{S12}. Therefore, bagging can be used to improve the $k$NN procedure in terms of both accuracy and stability when $d>2$. Furthermore, the CIS ratio of OWNN over BNN is less than $1$ for all $d$, but quickly converges to $1$ as $d$ increases. This implies that although the BNN procedure is asymptotically less stable than the OWNN procedure, their difference vanishes as $d$ increases.
\subsection{Comparisons between SNN and OWNN}\label{sec:regcis}
Corollary~\ref{cor:snn_ownn} in Section \ref{sec:ratesnn} implies that OWNN and SNN have the same convergence rates of regret and CIS (note that OWNN is a special case of SNN). Hence, it is of more interest to compare their relative magnitude. The asymptotic comparisons between SNN and OWNN are characterized in Corollary \ref{cor:snn_ownn_ratio}.
\begin{corollary}
\label{cor:snn_ownn_ratio}
Under Assumptions (A1)-(A4) defined in Appendix \ref{sec:assumptions} and the assumption that $B_2$ defined in Appendix \ref{sec:defwnb} is positive, we have, as $n\rightarrow \infty$,
\begin{eqnarray*}
\frac{\textrm{Regret}(\textrm{SNN})}{\textrm{Regret}(\textrm{OWNN})} &\longrightarrow& \Big\{\frac{B_1}{\lambda B_2}\Big\}^{d/(d+4)}\Big\{\frac{4+d\lambda B_2/B_1}{4+d}\Big\},\\
\frac{\textrm{CIS}(\textrm{SNN})}{\textrm{CIS}(\textrm{OWNN})} &\longrightarrow& \Big\{\frac{B_1}{\lambda B_2}\Big\}^{d/(2(d+4))}, \nonumber
\end{eqnarray*}
where constants $B_1$ and $B_2$ are defined in Appendix \ref{sec:defwnb}.
\end{corollary}
The second formula in Corollary~\ref{cor:snn_ownn_ratio} suggests that as $\lambda$ increases, the SNN classifier becomes more and more stable. In Corollary~\ref{cor:snn_ownn_ratio}, both ratios of the SNN procedure over the OWNN procedure depend on $\lambda$, and two unknown constants $B_1$ and $B_2$. Since $\lambda = (B_1+\lambda_0B_3^2)/B_2$ in $(\ref{formal})$ and $B_3=4B_1/\sqrt{\pi}$ in $(\ref{eq:asy_CIS})$, we further have the following ratios,
\begin{eqnarray}
\frac{\textrm{Regret}(\textrm{SNN})}{\textrm{Regret}(\textrm{OWNN})} &\longrightarrow& \Big\{\frac{1}{1 + 16B_1\lambda_0/\pi}\Big\}^{d/(d+4)}\Big\{\frac{4+d(1+16B_1\lambda_0/\pi)}{4+d}\Big\}, \label{snn_ownn_regret}\\
\frac{\textrm{CIS}(\textrm{SNN})}{\textrm{CIS}(\textrm{OWNN})} &\longrightarrow& \Big\{\frac{1}{1 + 16B_1\lambda_0/\pi}\Big\}^{d/(2(d+4))}. \label{snn_ownn_cis}
\end{eqnarray}
For any $\lambda_0>0$, SNN has an improvement in CIS over the OWNN. As a mere illustration, we consider the case that the regret and the squared CIS are given equal weight, that is, $\lambda_0=1$. In this case, the ratios in $(\ref{snn_ownn_regret})$ and $(\ref{snn_ownn_cis})$ only depend on $B_1$ and $d$.
\begin{figure}[h!]
\begin{center}
\includegraphics[scale=0.55]{regretratio_3d_new}
\includegraphics[scale=0.55]{cisratio_3d_new}
\caption{\label{ratio_po} \footnotesize Regret ratio and CIS ratio of SNN over OWNN as functions of $B_1$ and $d$. The darker the color, the larger the value. }
\end{center}
\end{figure}
Figure \ref{ratio_po} shows 3D plots of these two ratios as functions of $B_1$ and $d$. As expected, the CIS of the SNN procedure is universally smaller than OWNN (ratios less than 1 on the right panel), while the OWNN procedure has a smaller regret (ratios greater than 1 on the left panel). For a fixed $B_1$, as the dimension $d$ increases, the regret of SNN approaches that of OWNN, while the advantage of SNN in terms of CIS grows. For a fixed dimension $d$, as $B_1$ increases, the regret ratio between SNN and OWNN gets larger, but the CIS advantage of SNN also grows. According to the definition of $B_1$ in Appendix \ref{sec:defwnb}, a great value of $B_1$ indicates a harder problem for classification; see the discussion after Theorem 1 of \cite{S12}.
\begin{figure}[b!]
\begin{center}\vspace{-2em}
\includegraphics[scale=0.55]{loggain_3d_new}\vspace{-2em}
\caption{\footnotesize Logarithm of relative gain of SNN over OWNN as a function of $B_1$ and $d$ when $\lambda_0=1$. The grey (white) color represents the case where the logarithm of relative gain is greater (less) than $0$.}\label{loggain}
\end{center}
\end{figure}
Since SNN improves OWNN in CIS, but has a greater regret, it is of interest to know when the improvement of SNN in CIS is greater than its loss in regret. We thus consider the \textit{relative gain}, defined as the absolute ratio of the percentages of CIS reduction and regret increment, that is, $|\Delta \textrm{CIS}/\Delta \textrm{Regret}|$, where $\Delta \textrm{CIS} = [\textrm{CIS}(\textrm{SNN})-\textrm{CIS}(\textrm{OWNN})]/\textrm{CIS}(\textrm{OWNN})$ and $\Delta \textrm{Regret} = [\textrm{Regret}(\textrm{SNN})-\textrm{Regret}(\textrm{OWNN})]/\textrm{Regret}(\textrm{OWNN})$. As an illustration, when $\lambda_0=1$, we have the relative gain converges to $\left[1- (1+16B_1/\pi)^{-d/(2d+8)} \right]\left[ (1+16B_1/\pi)^{4/(d+4)} -1\right]^{-1}$. Figure \ref{loggain} shows the log(relative gain) as a function of $B_1$ and $d$. For most combinations of $B_1$ and $d$, the logarithm is greater than $0$ (shown in grey in Figure \ref{loggain}), indicating that SNN has an improvement in CIS greater than its loss in regret. In particular, when $B_1\le 0.2$, the log(relative gain) is positive for all $d$.
\section{Tuning Parameter Selection}\label{sec:algorithm}
To select the parameter $\lambda$ for the SNN classifier, we first identify a set of values for $\lambda$ whose corresponding (estimated) risks are among the smallest, and then choose from them an optimal one which has the minimal estimated CIS. Let $\widehat \phi^\lambda_{\mathcal D}$ denote an SNN classifier with parameter $\lambda$ trained from sample $\mathcal D$. Given a predetermined set of tuning parameter values $\Lambda = \{\lambda_1,\ldots,\lambda_K\}$, the tuning parameter $\widehat{\lambda}$ is selected using Algorithm 1 below, which involves estimating the CIS and risk in Steps 1--3 and a two-stage selection in Steps 4 and 5. \\
\indent {\it Algorithm 1:}\\
\indent {\it Step 1}. Randomly partition ${\cal D}=\{(X_i,Y_i),i=1,\ldots,n\}$ into five subsets $I_i$, $i=1,\cdots,5$.\\
\indent {\it Step 2}. For $i=1$, let $I_1$ be the test set and $I_2$, $I_3$, $I_4$ and $I_5$ be training sets. Obtain predicted labels from $\widehat{\phi}_{I_2\cup I_3}^{\lambda}(X_j)$ and $\widehat{\phi}_{I_4\cup I_5}^{\lambda}(X_j)$ respectively for each $X_j\in I_1$. Estimate the CIS and risk of the classifier with parameter $\lambda$ by
\begin{eqnarray*}
\widehat{\textrm{CIS}}_{i}(\lambda) &=& \frac{1}{|I_1|} \sum_{(X_j,Y_j)\in I_1} \ind{\widehat{\phi}_{I_2 \cup I_3}^{\lambda}(X_j) \ne \widehat{\phi}_{I_4\cup I_5}^{\lambda}(X_j)},\\
\widehat{\textrm{Risk}}_{i}(\lambda) &=& \frac{1}{2|I_1|} \sum_{(X_j,Y_j)\in I_1} \left\{ \ind{\widehat{\phi}_{I_2 \cup I_3}^{\lambda}(X_j) \ne Y_j} + \ind{\widehat{\phi}_{I_4\cup I_5}^{\lambda}(X_j)\ne Y_j}\right\}.
\end{eqnarray*}
\indent {\it Step 3}. Repeat {\it Step 2} for $i=2,\dots,5$ and estimate the CIS and risk, with $I_i$ being the test set and the rest being the training sets. Finally, the estimated CIS and risk are,
\begin{eqnarray*}
\widehat{\textrm{CIS}}(\lambda) = \frac{1}{5}\sum_{i=1}^{5} \widehat{\textrm{CIS}}_{i}(\lambda),~~
\widehat{\textrm{Risk}}(\lambda) = \frac{1}{5}\sum_{i=1}^{5} \widehat{\textrm{Risk}}_{i}(\lambda).
\end{eqnarray*}
\indent {\it Step 4}. Perform {\it Step 2} and {\it Step 3} for each $\lambda_k \in \Lambda$. Denote the set of tuning parameters with top accuracy as
$$
{\cal A} := \{ \lambda: \widehat{\textrm{Risk}}(\lambda) \textrm{~is less than the 10th percentile of~} \widehat{\textrm{Risk}}(\lambda_k),~k=1,\ldots,K\}.
$$
\indent {\it Step 5}. Output the optimal tuning parameter $\widehat{\lambda}$ as
$$
\widehat{\lambda} = \mathop{\rm argmin}_{\lambda \in {\cal A}} \widehat{\textrm{CIS}}(\lambda).
$$
In our experiments, the predetermined set of tuning parameters $\Lambda$ is of size $100$. In Step 1, the sample sizes of the subsets $I_i$ are chosen to be roughly equal. In Step 4, the threshold $10\%$ reflects how the set of the most accurate classifiers is defined. Based on our limited experiments, the final experimental result is very robust to the choice of this threshold level within a suitable range.
Compared with the tuning method for the $k$NN classifier, which minimizes the estimated risk only, Algorithm 1 requires additional estimation of the CIS. However, the estimation of the CIS is concurrently conducted with the estimation of the risk in Step 2. Therefore, the complexity of tuning for our SNN classifier is at the same order as that for $k$NN. As will be seen in the numerical experiments below, the additional effort on estimating the CIS leads to improvement over existing nearest neighbor methods in both accuracy and stability.
\section{Numerical Studies}\label{sec:exp}
We first verify our theoretical findings using an example, and then illustrate the improvements of the SNN classifier over existing nearest neighbor classifiers based on simulations and real examples.
\subsection{Validation of Asymptotically Equivalent Forms}\label{sec:validation}
This subsection aims to support the asymptotically equivalent forms of CIS derived in Theorem \ref{thm:CIS} and the CIS and regret ratios in Corollary \ref{cor:snn_ownn_ratio}. We focus on a multivariate Gaussian example in which regret and CIS have explicit expressions.
Assume that the underlying distributions of both classes are $P_1\sim N(0_2,\mathbb{I}_2)$ and $P_2\sim N(1_2,\mathbb{I}_2)$ and the prior class probability $\pi_1=1/3$. We choose ${\cal R}=[-2,3]^2$, which covers at least $95\%$ probability of the sampling region, and set $n=50, 100, 200$ and $500$. In addition, a test set with 1000 observations was independently generated. The estimated risk and CIS were calculated based on $100$ replications. In this example, some calculus exercises lead to $B_1=0.1299$, $B_2=10.68$ and $B_3=0.2931$. According to Proposition \ref{thm:regret}, Theorems \ref{thm:CIS} and \ref{thm:optimal}, we obtain that
\begin{align}
\textrm{Regret}(\textrm{SNN}) &= 0.1732{(k^*)}^{-1} + 4.7467 (k^*)^2n^{-2} \label{regret_sce1}\\
\textrm{CIS}(\textrm{SNN}) &= 0.3385(k^*)^{-1/2}, \label{CIS_sce2}
\end{align}
with $k^*= \lfloor 1.5^{1/3} \lambda^{1/3} n^{2/3}\rfloor$. For a mere illustration, we choose $\lambda=(B_1+B_3^2)/B_2$, which corresponds to $\lambda_0=1$. So we have $k^*=\lfloor 0.3118 n^{2/3}\rfloor.$
Similarly, the asymptotic regret and CIS of OWNN are (\ref{regret_sce1}) and (\ref{CIS_sce2}) with $k^*= \lfloor 0.2633 n^{2/3}\rfloor$ due to $(2.4)$ in \citet{S12}.
\begin{figure}[htb]
\begin{center}
\includegraphics[scale=0.6]{cis_ownn_new.pdf}
\includegraphics[scale=0.6]{cis_snn_new.pdf}
\caption{\label{fig:CIS_sce1} \footnotesize Asymptotic CIS (red curve) and estimated CIS (box plots over 100 simulations) for OWNN (left) and SNN (right) procedures. These plots show that the estimated CIS converges to its asymptotic equivalent value as $n$ increases.}
\end{center}
\end{figure}
In Figures \ref{fig:CIS_sce1}, we plot the asymptotic CIS of the SNN and OWNN classifiers computed using the above formulae, shown as red curves, along with the estimated CIS based on the simulated data, shown as the box plots over 100 replications. As the sample size $n$ increases, the estimated CIS approximates its asymptotic value very well. For example, when $n=500$, the asymptotic CIS of the SNN (OWNN) classifier is 0.078 (0.085) while the estimated CIS is 0.079 (0.086).
\begin{figure}[!b]
\begin{center}
\includegraphics[scale=0.6]{error_ownn_new.pdf}
\includegraphics[scale=0.6]{error_snn_new.pdf}\vspace{-1em}
\caption{\label{fig:error_sce1} \footnotesize Asymptotic risk (regret + the Bayes risk; red curves) and estimated risk (black box plots) for OWNN (left) and SNN procedures (right). The blue horizontal line indicates the Bayes risk, 0.215. These plots show that the estimated risk converges to its asymptotic version (and also the Bayes risk) as $n$ increases.}
\end{center}
\end{figure}
Similarly, in Figure \ref{fig:error_sce1}, we plot the asymptotic risk, that is, the asymptotic regret in (\ref{regret_sce1}) plus the true Bayes risk ($0.215$ in this example), for the SNN and OWNN classifiers, along with the estimated risk. Here we compute the Bayes risk by Monte Carlo integration. Again the difference of the estimated risk and asymptotic risk decreases as the sample size grows.
Furthermore, according to $(\ref{snn_ownn_cis})$, the asymptotic CIS ratio of the SNN classifier over the OWNN classifier is $0.9189$ in this example, and the empirically estimated CIS ratios are $0.6646$, $0.9114$, $0.8940$ and $0.9219$, for $n=50,100,200,500$. This indicates that the estimated CIS ratio converges to its asymptotic value as $n$ increases. However, by $(\ref{snn_ownn_regret})$, the asymptotic regret ratio of the SNN classifier over the OWNN classifier is $1.0305$, while the estimated ones are $1.0224$, $1.1493$, $0.3097$ and $0.1136$, for $n=50,100,200,500$. It appears that the estimated regret ratio matches with its asymptotic value for small sample size, but they differ for large $n$. This may be caused by the fact that the classification errors are very close to Bayes risk for large $n$ and hence the estimated regret ratio has a numerical issue. For example, when $n=500$, the average errors of the SNN classifier and the OWNN classifier are $0.2152$ and $0.2161$, respectively, while the Bayes risk is $0.215$ (see Figure \ref{fig:error_sce1}). A similar issue was previously reported in \citet{S12}.
\subsection{Simulations}
\label{sec:simulation}
In this section, we compare SNN with the $k$NN, OWNN and BNN classifiers. The parameter $k$ in $k$NN was tuned from $100$ equally spaced grid points from 5 to $n/2$. For a fair comparison, in the SNN classifier, the parameter $\lambda$ was tuned so that the corresponding parameter $k^*$ (see Theorem~\ref{thm:optimal}) were equally spaced and fell into the same range roughly.
In Simulation 1, we assumed that the two classes were from $P_1\sim N(0_d,\mathbb{I}_d)$ and $P_2\sim N(\mu_d,\mathbb{I}_d)$ with the prior probability $\pi_1=1/3$ and dimension $d$. We set sample size $n=200$ and chose $\mu$ such that the resulting $B_1$ was fixed as $0.1$ for different $d$. Specifically, in Supplementary \ref{proof53} we show that
\begin{equation}
B_1 = \frac{\sqrt{2\pi}}{3\pi\mu d}\exp\left(-\frac{( \mu d/2-\ln 2/\mu)^2}{2d}\right). \label{B1}
\end{equation}
Hence, we set $\mu=2.076$, $1.205$, $0.659$, $0.314$, $0.208$ for $d=1,2,4,8$ and $10$, respectively.
In Simulation 2, the training data set were generated by setting $n=200$, $d=2$ or $5$, $P_1\sim 0.5N(0_d,\mathbb{I}_d)+0.5 N(3_d,2\mathbb{I}_d)$, $P_2\sim 0.5N(1.5_d,\mathbb{I}_d)+0.5 N(4.5_d,2\mathbb{I}_d)$, and $\pi_1=1/2$ or $1/3$.
Simulation 3 has the same setting as Simulation 2, except that $P_1\sim 0.5N(0_d,\Sigma)+0.5 N(3_d,2\Sigma)$ and $P_2\sim 0.5N(1.5_d,\Sigma)+0.5 N(4.5_d,2\Sigma)$, where $\Sigma$ is the Toeplitz matrix whose $j$th entry of the first row is $0.6^{j-1}$.
\begin{figure}[!t]
\begin{center}
\includegraphics[scale=0.5]{sim1_error_new}
\includegraphics[scale=0.5]{sim1_cis_new}
\caption{\label{figure_sim1} \footnotesize Average test errors and CIS's (with standard error bar marked) of the $k$NN, BNN, OWNN, and SNN methods in Simulation 1. The $x$-axis indicates different settings with various dimensions. Within each setting, the four methods are horizontally lined up (from the left are $k$NN, BNN, OWNN, and SNN).}
\end{center}
\end{figure}
Simulation 1 is a relatively easy classification problem. Simulation 2 examines the bimodal effect and Simulation 3 combines bimodality with dependence between variables. In each simulation setting, a test data set of size $1000$ is independently generated and the average classification error and average estimated CIS for the test set are reported over $100$ replications. To estimate the CIS, for each replication, we build two classifiers based on the randomly divided training data, and then estimate CIS by the average disagreement of these two classifiers on the test data.
Figure \ref{figure_sim1} shows the average error (on the left) and CIS (on the right) for Simulation 1. As a first impression, the test error is similar among different classification methods, while the CIS differs a lot. In terms of the stability, SNN always has the smallest CIS; in particular, as $d$ increases, the improvement of SNN over all other procedures becomes even larger. This agrees with the asymptotic findings in Section \ref{sec:regcis}. For example, when $d=10$, all the $k$NN, BNN, and OWNN procedures are at least five times more unstable than SNN. In terms of accuracy, SNN obtains the minimal test errors in all five scenarios, although the improvement in the accuracy is not significant when $d=1,~2$ or $4$. This result suggests that although SNN is asymptotically less accurate than OWNN in theory, the actual empirical difference in the test error is often ignorable.
\begin{figure}[!t]
\begin{center}
\includegraphics[scale=0.5]{sim2_error_new}
\includegraphics[scale=0.5]{sim2_cis_new}
\caption{\label{figure_sim2} \footnotesize Average test errors and CIS's (with standard error bar marked) of the $k$NN, BNN, OWNN, and SNN methods in Simulation 2. The ticks on the $x$-axis indicate the dimensions and prior class probability $\pi$ for different settings. Within each setting, the four methods are horizontally lined up (from the left are $k$NN, BNN, OWNN, and SNN).}
\end{center}
\end{figure}
\begin{figure}[t!]
\begin{center}
\includegraphics[scale=0.5]{sim3_error_new}
\includegraphics[scale=0.5]{sim3_cis_new}
\caption{\label{figure_sim3} \footnotesize Average test errors and CIS's (with standard error bar marked) of the $k$NN, BNN, OWNN, and SNN methods in Simulation 3. The ticks on the $x$-axis indicate the dimensions and prior class probability $\pi$ for different settings. Within each setting, the four methods are horizontally lined up (from the left are $k$NN, BNN, OWNN, and SNN).}
\end{center}
\end{figure}
Figures \ref{figure_sim2} and \ref{figure_sim3} summarize the results for Simulations 2 and 3. Similarly, in general, the difference in CIS is much obvious than the difference in the error. The SNN procedure obtains the minimal CIS in all 8 cases. Interestingly, the improvements are significant in all the four cases when $\pi_1 = 1/3$. Moreover, among 3 out of the 8 cases, our SNN achieves the smallest test errors and the improvements are significant. Even in cases where the error is not the smallest, the accuracy of SNN is close to the best classifier.
\subsection{Real Examples}
\label{real}
\begin{figure}[b!]
\begin{center}
\includegraphics[scale=0.5]{real_error_new}
\includegraphics[scale=0.5]{real_cis_new}
\caption{\label{figure_real} \footnotesize Average test errors and CIS's (with standard error bar marked) of the $k$NN, BNN, OWNN and SNN methods for four data examples. The ticks on the $x$-axis indicate the names of the examples. Within each example, the four methods are horizontally lined up (from the left are $k$NN, BNN, OWNN, and SNN).}
\end{center}
\end{figure}
We extend the comparison to four real data sets publicly available in the UCI Machine Learning Repository \citep{BL13}. The first data set is the breast cancer data set ($breast$) collected by \citet{WM90}. There are 683 samples and 10 experimental measurement variables. The binary class label indicates whether the sample is benign or malignant. These 683 samples arrived periodically. In total, there are 8 groups of samples which reflect the chronological order of the data. A good classification procedure is expected to produce a stable classifier across these groups of samples. The second data set is the credit approval data set ($credit$). It consists of $690$ credit card applications and each application has $14$ attributes reflecting the user information. The binary class label refers to whether the application is positive or negative. The third data set is the haberman's survival data set ($haberman$) which contains $306$ cases from study conducted on the survival of patients who had undergone surgery for breast cancer. It has three attributes, age, patient's year of operation, and number of positive axillary nodes detected. The response variable indicates the survival status: either the patient survived 5 years or longer or the patient died within 5 years. The last data set is the SPECT heart data set ($spect$) which describes the diagnosing of cardiac Single Proton Emission Computed Tomography (SPECT) images. Each of the $267$ image sets (patients) had $22$ binary feature patterns and was classified into two classes: normal and abnormal.
We randomly split each data set into training and test sets with the equal size. The same tuning procedure as in the simulation is applied here. We compute the test error and the estimated CIS on the test set. The procedure is repeated 100 times and the average error and CIS are reported in Figure \ref{figure_real}.
Similar to the simulation results, the SNN procedure obtains the minimal CIS in all four real data sets and the improvements in CIS are significant. The errors of OWNN and SNN have no significant difference, although OWNN is theoretically better in accuracy. These real experiments further illustrate that, with almost the same classification accuracy, our SNN procedure can achieve a significant improvement in the stability, which promotes the reproducibility.
\section{Conclusion}
\label{discussion}
Stability is an important and desirable property of a statistical procedure. It provides a foundation for the reproducibility, and reflects the credibility of those who use the procedure. To our best knowledge, our work is the first to propose a measure to quantify classification instability. The proposed SNN classification procedure enjoys increased classification stability with comparable classification accuracy to OWNN.
For classification problems, the classification accuracy is a primary concern, while stability is secondary. In many real cases, however, different classifiers may enjoy a comparable classification accuracy, and a classifier with a better stability stands out. The observation that our method can improve stability while maintaining the similar accuracy suggests that there may exist much more room for improving stability than for improving accuracy. This may be explained by the faster convergence rate of the regret than that of the CIS (Theorem \ref{thm:upperCISsnn}).
In theory, our SNN is shown to achieve the minimax optimal convergence rate in regret and a sharp convergence rate in CIS. Extensive experiments illustrate that SNN attains a significant improvement of stability over existing nearest neighbor classifiers, and sometimes even improves the accuracy. We implement the algorithm in a publicly available R package \texttt{snn}.
Our proposed SNN method is motivated by an asymptotic expansion of the CIS. Such a nice property may not exist for other more general classification methods. Hence, it is unclear how the stabilization idea can be carried over to other classifiers in a similar manner. That being said, the CIS measure can be used as a criterion for tuning parameter selection. There exists work in the literature which uses variable selection stability to select tuning parameter \citep{SWF13}. Classification stability and variable selection stability complement each other and can provide a comprehensive description of the reliability of a statistical procedure.
For simplicity, we focus on the binary classification in this article. The generalization of the SNN classifier to multi-category classification problems \citep{LLW04, LS06, LY11} is an interesting topic to pursue in the future. Moreover, stability for the high-dimensional, low-sample size data is another important topic. Furthermore, in analyzing a big data set, a popular scheme is divide-and-conquer. It is an interesting research question on how to divide the data and choose the parameter wisely to ensure the optimal stability of a combined classifier.
\section*{Appendices}
\setcounter{subsection}{0}
\renewcommand{\thesubsection}{A.\Roman{subsection}}
\setcounter{equation}{0}
\renewcommand{\theequation}{A.\arabic{equation}}
\setcounter{lemma}{0}
\renewcommand{\thelemma}{A.\arabic{lemma}}
\subsection{Assumptions (A1) - (A4)}
\label{sec:assumptions}
For a smooth function $g$, we denote $\dot{g}(x)$ as its gradient vector at $x$. We assume the following conditions through all the article.
(A1) The set ${\cal R}\subset \mathbb R^d$ is a compact $d$-dimensional manifold with boundary $\partial{\cal R}$.
(A2) The set ${\cal S}=\{x\in {\cal R}: \eta(x)=1/2\}$ is nonempty. There exists an open subset $U_0$ of ${\mathbb R}^d$ which contains ${\cal S}$ such that: (i) $\eta$ is continuous on $U\backslash U_0$ with $U$ an open set containing ${\cal R}$; (ii) the restriction of the conditional distributions of $X$, $P_1$ and $P_2$, to $U_0$ are absolutely continuous with respect to Lebesgue measure, with twice continuously differentiable Randon-Nikodym derivatives $f_1$ and $f_2$.
(A3) There exists $\rho >0$ such that $\int_{{\mathbb R}^d}\|x\|^{\rho} d\bar{P}(x) < \infty$. Moreover, for sufficiently small $\delta>0$, $\inf_{x\in {\cal R}}\bar{P}(B_{\delta}(x))/(a_d\delta^d) \ge C_3 >0$, where $a_d=\pi^{d/2}/\Gamma(1+d/2)$, $\Gamma(\cdot)$ is gamma function, and $C_3$ is a constant independent of $\delta$.
(A4) For all $x\in {\cal S}$, we have $\dot{\eta}(x)\ne 0$, and for all $x\in {\cal S}\cap \partial{\cal R}$, we have $\dot{\partial \eta}(x)\ne 0$, where $\partial \eta$ is the restriction of $\eta$ to $\partial {\cal R}$. \hfill $\blacksquare$
\begin{remark}
Assumptions (A1)--(A4) have also been employed to show the asymptotic expansion of the regret of the $k$NN classifier \citep{HPS08}. The condition $\dot{\eta}(x)\ne 0$ in (A4) is equivalent to the margin condition with $\alpha=1$; see (2.1) in \citet{S12}. Furthermore, these assumptions ensure that $\bar{f}(x_0)$ and $\dot{\eta}(x_0)$ are bounded away from zero and infinity on ${\cal S}$.
\end{remark}
\subsection{Definitions of $a(x)$, $B_1$, $B_2$, and $W_{n,\beta}$}\label{sec:defwnb}
For a smooth function $g$: $\mathbb{R}^d\rightarrow \mathbb{R}$, let $g_j(x)$ its $j$th partial derivative at $x$, $\ddot{g}(x)$ the Hessian matrix at $x$, and $g_{jk}(x)$ the $(j,k)$th element of $\ddot{g}(x)$. Let $c_{j,d}=\int_{v:\|v\|\le 1} v_j^2 dv$. Define
$$
a(x)=\sum_{j=1}^d \frac{c_{j,d}\{\eta_{j}(x)\bar{f}_j(x) + 1/2 \eta_{jj}(x)\bar{f}(x)\}}{a_d^{1+2/d} \bar{f}(x)^{1+2/d}}.
$$
We further define two distribution-related constants
\begin{eqnarray*}
B_1 = \int_{\cal S} \frac{\bar{f}(x)}{4\|\dot{\eta}(x)\|} d \textrm{Vol}^{d-1}(x),\quad B_2 = \int_{\cal S} \frac{\bar{f}(x)}{\|\dot{\eta}(x)\|} a(x)^2 d \textrm{Vol}^{d-1}(x),
\end{eqnarray*}
where $\textrm{Vol}^{d-1}$ is the natural $(d-1)$-dimensional volume measure that ${\cal S}$ inherits with ${\cal S}$ defined in Appendix \ref{sec:assumptions}. Based on Assumptions (A1)-(A4) in Appendix \ref{sec:assumptions}, $B_1$ and $B_2$ are finite with $B_1>0$ and $B_2\ge 0$, where $B_2=0$ only when $a(x)$ equals zero on ${\cal S}$.
In addition, for $\beta>0$, we denote $W_{n,\beta}$ as the set of $\boldsymbol{w}_n$ satisfying (w.1)--(w.5).
(w.1) $\sum_{i=1}^n w_{ni}^2 \le n^{-\beta}$,
(w.2) $n^{-4/d}(\sum_{i=1}^n\alpha_iw_{ni})^2\le n^{-\beta}$, where $\alpha_i=i^{1+\frac{2}{d}}-(i-1)^{1+\frac{2}{d}}$,
(w.3) $n^{2/d}\sum_{i=k_2+1}^n w_{ni}/\sum_{i=1}^n \alpha_iw_{ni}\le 1/\log n$ with $k_2=\lfloor n^{1-\beta} \rfloor$,
(w.4) $\sum_{i=k_2+1}^n w_{ni}^2/\sum_{i=1}^n w_{ni}^2 \le 1/\log n$,
(w.5) $\sum_{i=1}^n w_{ni}^3/(\sum_{i=1}^n w_{ni}^2)^{3/2} \le 1/\log n$.
For the $k$NN classifier with $w_{ni}=k^{-1}\ind{1\le i\le k}$, (w.1)--(w.5) reduce to $\max(n^{\beta}, (\log n)^2) \le k \le \min(n^{(1-\beta d /4)}, n^{1-\beta})$. See \citet{S12} for a detailed discussion of $W_{n,\beta}$. \hfill $\blacksquare$
\subsection{Proof of Theorem \ref{thm:CIS}}
\label{sec:proofcis}
Note that $\textrm{CIS}(\textrm{WNN})= {\mathbb P}_{{\cal D}_1,{\cal D}_2,X}\Big(\widehat{\phi}_{n1}^{\boldsymbol{w}_n}(X) \ne \widehat{\phi}_{n2}^{\boldsymbol{w}_n}(X)\Big)$ can be expressed in the following way.
\begin{eqnarray*}
&&\textrm{CIS}(\textrm{WNN}) \\
&=& {\mathbb E}_{X} \Big[ {\mathbb P}_{{\cal D}_1,{\cal D}_2}\Big(\widehat{\phi}_{{\cal D}_1}^{\boldsymbol{w}_n}(X) \ne \widehat{\phi}_{{\cal D}_2}^{\boldsymbol{w}_n}(X)\Big|X\Big) \Big]\\
&=& {\mathbb E}_{X} \Big[ {\mathbb P}_{{\cal D}_1,{\cal D}_2}\Big(\widehat{\phi}_{{\cal D}_1}^{\boldsymbol{w}_n}(X)=1, \widehat{\phi}_{{\cal D}_2}^{\boldsymbol{w}_n}(X) =2 \Big|X \Big)\Big] + {\mathbb E}_{X} \Big[ {\mathbb P}_{{\cal D}_1,{\cal D}_2}\Big(\widehat{\phi}_{{\cal D}_1}^{\boldsymbol{w}_n}(X)=2, \widehat{\phi}_{{\cal D}_2}^{\boldsymbol{w}_n}(X) =1 \Big|X\Big)\Big] \\
&=& {\mathbb E}_{X} \Big[ 2 {\mathbb P}_{{\cal D}_1}\Big(\widehat{\phi}_{{\cal D}_1}^{\boldsymbol{w}_n}(X)=1|X\Big) \Big(1- {\mathbb P}_{{\cal D}_1}\Big(\widehat{\phi}_{{\cal D}_1}^{\boldsymbol{w}_n}(X)=1|X\Big)\Big) \Big],
\end{eqnarray*}
where the last equality is valid because ${\cal D}_1$ and ${\cal D}_2$ are i.i.d. samples. Without loss of generality, we consider a generic sample ${\cal D}=\{(X_i,Y_i),i=1,\ldots,n\}$. Given $X=x$, we define $(X_{(i)},Y_{(i)})$ such that $\|X_{(1)}-x\|\le \|X_{(2)}-x\|\le \ldots \le \|X_{(n)}-x\|$ with $\|\cdot\|$ the Euclidean norm. Denote the estimated regression function $S_n(x) = \sum_{i=1}^n w_{ni}\ind{Y_{(i)}=1}$. We have
\begin{align*}
{\mathbb E}_{X} \Big[ {\mathbb P}\Big(\widehat{\phi}_{{\cal D}}^{\boldsymbol{w}_n}(X)=1|X\Big)\Big] &= \int_{{\cal R}} {\mathbb P}\Big(S_n(x) \ge 1/2 \Big) d \bar{P}(x),\\
{\mathbb E}_{X} \Big[{\mathbb P}^2\Big(\widehat{\phi}_{{\cal D}}^{\boldsymbol{w}_n}(X)=1|X\Big)\Big] &= \int_{{\cal R}} {\mathbb P}^2\Big(S_n(x) \ge 1/2 \Big) d \bar{P}(x),
\end{align*}
where $\bar{P}(x)$ is the marginal distribution of $X$. For the sake of simplicity, $\mathbb P$ denotes the probability with respect to $\cal D$. Hence, CIS satisfies
\begin{align*}
\textrm{CIS}(\textrm{WNN})/2 &= \int_{{\cal R}} {\mathbb P}(S_n(x) \ge 1/2 ) \Big(1- {\mathbb P}(S_n(x) \ge 1/2)\Big)d \bar{P}(x)\\
&= \int_{{\cal R}} \left\{ {\mathbb P}(S_n(x) < 1/2 ) - \ind{\eta(x)<1/2}\right\} d\bar{P}(x) \nonumber\\
&~~~ - \int_{{\cal R}} \left\{ {\mathbb P}^2(S_n(x) < 1/2) - \ind{\eta(x)<1/2} \right\}d \bar{P}(x)
\end{align*}
Denote the boundary ${\cal S}=\{x\in {\cal R}: \eta(x)=1/2\}$. For $\epsilon>0$, let ${\cal S}^{\epsilon\epsilon} = \{x\in {\mathbb R}^d: \eta(x)=1/2 ~\textrm{and}~ \textrm{dist}(x, {\cal S})<\epsilon \}$, where $\textrm{dist}(x, {\cal S})=\inf_{x_0\in {\cal S}} \|x-x_0\|$. We will focus on the set
$$
{\cal S}^{\epsilon} = \left\{x_0 + t\frac{\dot{\eta}(x_0)}{\|\dot{\eta}(x_0)\|}: x_0 \in {\cal S}^{\epsilon\epsilon}, |t| < \epsilon \right\}.
$$
Let $\mu_n(x)={\mathbb E}\{S_n(x)\}$, $\sigma_n^2(x)=\textrm{Var}\{S_n(x)\}$, and $\epsilon_n=n^{-\beta d/4}$. Denote $s_n^2=\sum_{i=1}^n w_{ni}^2$ and $t_n=n^{-2/d}\sum_{i=1}^n \alpha_i w_{ni}$. \citet{S12} showed that, uniformly for $\boldsymbol{w}_n\in W_{n,\beta}$,
\begin{eqnarray}
\sup_{x\in {\cal S}^{\epsilon_n}} |\mu_n(x) - \eta(x) - a(x)t_n| &=& o(t_n),\label{step1:tn}\\
\sup_{x\in {\cal S}^{\epsilon_n}} \left|\sigma_n^2(x)-\frac{1}{4}s_n^2\right| &=& o(s_n^2). \label{step1:sn}
\end{eqnarray}
We organize our proof in three steps. In Step 1, we focus on analyzing on the set ${\cal R} \cap {\cal S}^{\epsilon_n} $; in Step 2, we focus on the complement set ${\cal R} \backslash {\cal S}^{\epsilon_n} $; Step 3 combines the results and applies a normal approximation to yield the final conclusion.
{\it Step 1}: For $x_0\in {\cal S}$ and $t\in {\mathbb R}$, denote $x_0^t=x_0+t \dot{\eta}(x_0)/\|\dot{\eta}(x_0)\|$. Denote $\bar{f}=\pi_1 f_1 + (1-\pi_1) f_2$ as the Radon-Nikodym derivative with respect to Lebesgue measure of the restriction of $\bar{P}$ to ${\cal S}^{\epsilon_n}$ for large $n$. We need to show that, uniformly for $\boldsymbol{w}_n\in W_{n,\beta}$,
\begin{eqnarray}
&&\int_{{\cal R}\cap{{\cal S}^{\epsilon_n}}} \left\{ {\mathbb P}(S_n(x) < 1/2) - \ind{\eta(x)<1/2} \right\} d \bar{P}(x) = \nonumber\\
&&~~~\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0)\{1+o(1)\};\label{tube}\\
&&\int_{{\cal R}\cap{{\cal S}^{\epsilon_n}}} \left\{ {\mathbb P}^2(S_n(x) < 1/2) - \ind{\eta(x)<1/2} \right\} d \bar{P}(x) = \nonumber\\
&&~~~\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}^2\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0)\{1+o(1)\}.\label{tube2}
\end{eqnarray}
According to \citet{S12}, for large $n$, we define the map $\phi(x_0,t\frac{\dot{\eta}(x_0)}{\|\dot{\eta}(x_0)\|})=x_0^t$, and note that
$$
\det \dot{\phi}\Big(x_0,t\frac{\dot{\eta}(x_0)}{\|\dot{\eta}(x_0)\|}\Big)dt d\textrm{Vol}^{d-1}(x_0) = \{1+o(1)\}dt d\textrm{Vol}^{d-1}(x_0),
$$
uniformly in $(x_0, t \dot{\eta}(x_0)/\|\dot{\eta}(x_0)\|)$ for $x_0\in {\cal S}$ and $|t|<\epsilon_n$, where $\det$ is the determinant. Then the theory of integration on manifolds \citep{G04} implies that, uniformly for $\boldsymbol{w}_n\in W_{n,\beta}$,
\begin{eqnarray}
&&\int_{{\cal S}^{\epsilon_n}} \left\{ {\mathbb P}(S_n(x) < 1/2) - \ind{\eta(x)<1/2} \right\} d \bar{P}(x) = \nonumber\\
&&\int_{{\cal S}^{\epsilon_n\epsilon_n}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0)\{1+o(1)\}.\nonumber
\end{eqnarray}
Furthermore, we can replace ${\cal S}^{\epsilon_n}$ with ${\cal R}\cap{{\cal S}^{\epsilon_n}}$ since ${\cal S}^{\epsilon_n}\backslash {\cal R}\subseteq \{x\in \mathbb{R}^d: \textrm{dist}(x,\partial {\cal S})<\epsilon_n\}$ and the latter has volume $O(\epsilon_n^2)$ by Weyl's tube formula \citep{G04}. Similarly, we can safely replace ${\cal S}^{\epsilon_n\epsilon_n}$ with ${\cal S}$. Therefore, $(\ref{tube})$ holds. Similar arguments imply $(\ref{tube2})$.
{\it Step 2}: Bound the contribution to CIS from ${\cal R}\backslash {\cal S}^{\epsilon_n}$. We show that, for all $M>0$,
\begin{eqnarray}
\sup_{\boldsymbol{w}_n\in W_{n,\beta}} \int_{{\cal R}\backslash {\cal S}^{\epsilon_n}} \left\{ {\mathbb P}\Big(S_n(x) < 1/2\Big) - \ind{\eta(x)<1/2} \right\} d \bar{P}(x) &=& O(n^{-M}),\label{bound1}\\
\sup_{\boldsymbol{w}_n\in W_{n,\beta}} \int_{{\cal R}\backslash {\cal S}^{\epsilon_n}} \left\{ {\mathbb P}^2\Big(S_n(x) < 1/2\Big) - \ind{\eta(x)<1/2} \right\} d \bar{P}(x) &=& O(n^{-M}).\label{bound2}
\end{eqnarray}
Here $(\ref{bound1})$ follows from the fact $|{\mathbb P}(S_n(x) < \frac{1}{2}) - \ind{\eta(x)<1/2}| = O(n^{-M})$ for all $M>0$, uniformly for $\boldsymbol{w}_n\in W_{n,\beta}$ and $x\in {\cal R}\backslash {\cal S}^{\epsilon_n}$ \citep{S12}. Furthermore, $(\ref{bound2})$ holds since
$$
\Big|{\mathbb P}^2\Big(S_n(x) < 1/2\Big) - \ind{\eta(x)<1/2}\Big|\le 2\Big|{\mathbb P}\Big(S_n(x) < 1/2\Big) - \ind{\eta(x)<1/2}\Big|.
$$
{\it Step 3}: In the end, we will show
\begin{eqnarray}
&&\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&&~~~-\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}^2\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&=& \frac{1}{2}B_3 s_n + o(s_n+t_n).
\end{eqnarray}
We first apply the nonuniform version of Berry-Esseen Theorem to approximate ${\mathbb P}(S_n(x_0^t) < 1/2)$. Let $Z_i= (w_{ni}\ind{Y_{(i)}=1} - w_{ni} \mathbb{E} [\ind{Y_{(i)}=1}])/\sigma_n(x)$ and $W=\sum_{i=1}^n Z_i$. Note that $\mathbb{E}(Z_i)=0$, $\textrm{Var}(Z_i)<\infty$, and $\textrm{Var}(W)=1$. Then the nonuniform Berry-Esseen Theorem \citep{B77} implies that
$$
\Big|\mathbb{P}(W\le y) - \Phi(y)\Big| \le \frac{M_1}{n^{1/2}( 1 + |y|^3)},
$$
where $\Phi$ is the standard normal distribution function and $M_1$ is a constant. Therefore,
\begin{equation}
\sup_{x_0\in {\cal S}}\sup_{t\in [-\epsilon_n,\epsilon_n]} \left|\mathbb{P}\Big(\frac{S_n(x_0^t)-\mu_n(x_0^t)}{\sigma_n(x_0^t)}\le y \Big) - \Phi(y) \right| \le \frac{M_1}{n^{1/2} (1 + |y|^3)}. \label{BE_thm}
\end{equation}
Thus, we have
\begin{eqnarray}
&&\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&&~~~~~~= \int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \bar{f}(x_0^t) \left\{\Phi\Big(\frac{1/2 - \mu_n(x_0^t)}{\sigma_n(x_0^t)} \Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) + o(s_n^2+t_n^2), \nonumber
\end{eqnarray}
where the remainder term $o(s_n^2+t_n^2)$ is due to $(\ref{BE_thm})$ by slightly modifying the proof of A.21 in \citet{S12}.
Furthermore, Taylor expansion leads to
$$
\bar{f}(x_0^t) = \bar{f}(x_0) + (\dot{\bar{f}}(x_0))^T\frac{\dot{\eta}(x_0)}{\|\dot{\eta}(x_0)\|} t + o(t).
$$
Therefore,
\begin{eqnarray}
&&\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) \label{exp1}\\
&=& \int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \bar{f}(x_0) \left\{\Phi\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0)\nonumber\\
&&+\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \frac{\dot{\bar{f}}(x_0)^T\dot{\eta}(x_0)t}{\|\dot{\eta}(x_0)\|} \left\{\Phi\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) + R_1, \nonumber
\end{eqnarray}
where
\begin{align*}
R_1 &= \int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \bar{f}(x_0) \left\{ \Phi\Big(\frac{1/2 - \mu_n(x_0^t)}{\sigma_n(x_0^t)}\Big)- \Phi\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big)\right\} dt d\textrm{Vol}^{d-1}(x_0)\\
&~~~~ + \int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \frac{\dot{\bar{f}}(x_0)^T\dot{\eta}(x_0)t}{\|\dot{\eta}(x_0)\|} \left\{ \Phi\Big(\frac{1/2 - \mu_n(x_0^t)}{\sigma_n(x_0^t)}\Big)- \Phi\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big)\right\} dt d\textrm{Vol}^{d-1}(x_0)\\
&~~~~ + o(s_n^2+t_n^2)\\
&\buildrel \Delta \over = R_{11}+R_{12}+o(s_n^2+t_n^2).
\end{align*}
Next we show $R_1=o(s_n+t_n)$. Denote $r_{x_0}=\frac{-a(x_0)t_n}{\|\dot{\eta}(x_0)s_n\|}$. According to $(\ref{step1:tn})$ and $(\ref{step1:sn})$, for a sufficiently small $\epsilon\in (0,\inf_{x_0\in {\cal S}} \|\dot{\eta}(x_0)\|)$ and large $n$, for all $\boldsymbol{w}_n\in W_{n,\beta}$, $x_0\in {\cal S}$ and $r\in[-\epsilon_n/s_n,\epsilon_n/s_n]$, \citet{S12} showed that
$$
\Big| \frac{1/2-\mu_n(x_0^{rs_n})}{\sigma_n(x_0^{rs_n})} -[-2\|\dot{\eta}(x_0)\|(r-r_{x_0})] \Big| \le \epsilon^2(|r|+t_n/s_n).
$$
In addition, when $|r-r_{x_0}|\le \epsilon t_n/s_n$,
\begin{eqnarray*}
\Big| \Phi\Big(\frac{1/2-\mu_n(x_0^{rs_n})}{\sigma_n(x_0^{rs_n})}\Big) -\Phi\Big(-2\|\dot{\eta}(x_0)\|(r-r_{x_0})\Big) \Big|\le 1
\end{eqnarray*}
and when $\epsilon t_n/s_n < |r| < t_n/s_n$,
\begin{eqnarray*}
\Big| \Phi\Big(\frac{1/2-\mu_n(x_0^{rs_n})}{\sigma_n(x_0^{rs_n})}\Big) -\Phi\Big(-2\|\dot{\eta}(x_0)\|(r-r_{x_0})\Big) \Big|\le \epsilon^2(|r|+t_n/s_n)\phi(\|\dot{\eta}(x_0)\||r-r_{x_0}|),
\end{eqnarray*}
where $\phi$ is the density function of standard normal distribution.
Therefore, we have
\begin{eqnarray}
|R_{11}| &\le & \int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \bar{f}(x_0) \left| \Phi\Big(\frac{1/2 - \mu_n(x_0^t)}{\sigma_n(x_0^t)}\Big)- \Phi\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big)\right| dt d\textrm{Vol}^{d-1}(x_0)\nonumber\\
&\le & \bar{f}(x_0) s_n \int_{|r-r_{x_0}|\le \epsilon t_n/s_n} dr + \bar{f}(x_0) s_n \epsilon^2 \int_{-\infty}^{\infty} (|r|+t_n/s_n)\phi(\|\dot{\eta}(x_0)\||r-r_{x_0}|) dr\nonumber\\
&\le& \epsilon (t_n + s_n) \label{R11}.
\end{eqnarray}
Similarly,
\begin{eqnarray*}
&&|R_{12}| \nonumber\\
&\le & \int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \frac{\dot{\bar{f}}(x_0)^T\dot{\eta}(x_0)t}{\|\dot{\eta}(x_0)\|} \left| \Phi\Big(\frac{1/2 - \mu_n(x_0^t)}{\sigma_n(x_0^t)}\Big)- \Phi\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big)\right| dt d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&\le & \bar{f}(x_0) \epsilon s_n^2 \int_{|r-r_{x_0}|\le \epsilon t_n/s_n} |r| dr + \bar{f}(x_0) s_n^2 \epsilon^2 \int_{-\infty}^{\infty} (|r|+t_n/s_n)\phi(\|\dot{\eta}(x_0)\||r-r_{x_0}|) dr \nonumber\\
&\le& \epsilon (t_n^2 + s_n^2).
\end{eqnarray*}
The inequality above, along with with $(\ref{R11})$, leads to $R_1=o(s_n+t_n)$.
By similar arguments, we have
\begin{align}
&\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} {\bar f}(x_0^t) \left\{{\mathbb P}^2\Big(S_n(x_0^t) < 1/2\Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) \label{exp2}\\
=& \int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \bar{f}(x_0) \left\{\Phi^2\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0)\nonumber\\
&+\int_{{\cal S}} \int_{-\epsilon_n}^{\epsilon_n} \frac{\dot{\bar{f}}(x_0)^T\dot{\eta}(x_0)t}{\|\dot{\eta}(x_0)\|} \left\{\Phi^2\Big(\frac{-2t\|\dot{\eta}(x_0)\|- 2a(x_0)t_n}{s_n} \Big) - \ind{t<0}\right\} dt d\textrm{Vol}^{d-1}(x_0)\nonumber\\
&+ o(s_n+t_n). \nonumber
\end{align}
Finally, after substituting $t=us_n/2$ in (\ref{exp1}) and (\ref{exp2}), we have, up to $o(s_n+t_n)$ difference,
\begin{eqnarray*}
&&\textrm{CIS}(\textrm{WNN})/2 \nonumber\\
&=&\frac{s_n}{2}\int_{{\cal S}} \int_{-\infty}^{\infty} \bar{f}(x_0) \left\{\Phi\Big(-\|\dot{\eta}(x_0)\|u-\frac{2a(x_0)t_n}{s_n} \Big) - \ind{u<0} \right\} du d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&&+ \frac{s_n^2}{4}\int_{{\cal S}} \int_{-\infty}^{\infty} \frac{(\dot{\bar{f}}(x_0))^T\dot{\eta}(x_0)}{\|\dot{\eta}(x_0)\|} u \left\{\Phi\Big(-\|\dot{\eta}(x_0)\|u-\frac{2a(x_0)t_n}{s_n} \Big) - \ind{u<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&&- \frac{s_n}{2}\int_{{\cal S}} \int_{-\infty}^{\infty} \bar{f}(x_0) \left\{\Phi^2\Big(-\|\dot{\eta}(x_0)\|u-\frac{2a(x_0)t_n}{s_n} \Big) - \ind{u<0}\right\} du d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&& - \frac{s_n^2}{4}\int_{{\cal S}} \int_{-\infty}^{\infty} \frac{(\dot{\bar{f}}(x_0))^T\dot{\eta}(x_0)}{\|\dot{\eta}(x_0)\|} u \left\{\Phi^2\Big(-\|\dot{\eta}(x_0)\|u-\frac{2a(x_0)t_n}{s_n} \Big) - \ind{u<0}\right\} dt d\textrm{Vol}^{d-1}(x_0) \nonumber\\
&=& I + II - III - IV.
\end{eqnarray*}
According to Lemma \ref{lemma}, we have
\begin{eqnarray*}
I - III &=& \left [\int_{\cal S} \frac{{\bar{f}}(x_0)}{2\sqrt{\pi}\|\dot{\eta}(x_0)\|} d\textrm{Vol}^{d-1}(x_0) \right] s_n = \frac{1}{2}B_3 s_n\\
II - IV &=& - \left [\int_{\cal S} \frac{(\dot{\bar{f}}(x_0))^T\dot{\eta}(x_0)a(x_0)}{2\sqrt{\pi}(\|\dot{\eta}(x_0)\|)^3} d\textrm{Vol}^{d-1}(x_0)\right] s_nt_n = \frac{1}{2}B_4s_nt_n.
\end{eqnarray*}
Therefore, the desirable result is obtained by noting that $B_4 s_nt_n = o(s_n+t_n)$. This concludes the proof of Theorem \ref{thm:CIS}. \hfill $\blacksquare$
\subsection{Proof of Theorem \ref{thm:optimal}} For any weight $\boldsymbol{w}_n$, the Lagrangian of $(\ref{formal})$ is
$$
L(\boldsymbol{w}_n)=\Big(\sum_{i=1}^n \frac{\alpha_i w_{ni}}{n^{2/d}}\Big)^2 + \lambda \sum_{i=1}^n w_{ni}^2 + \nu(\sum_{i=1}^{n}w_{ni}-1).
$$
Considering the constraint of nonnegative weights, we denote $k^{*}=\max\{i:w_{ni}^{*}>0\}$. Setting derivative of $L(\boldsymbol{w}_n)$ to be $0$, we have
\begin{equation}
\frac{\partial L(\boldsymbol{w}_n)}{\partial w_{ni}} = 2 n^{-4/d}\alpha_i\sum_{i=1}^{k^*} \alpha_i w_{ni} + 2\lambda w_{ni} + \nu =0. \label{derivative}
\end{equation}
Summing $(\ref{derivative})$ from 1 to $k^*$, and multiplying $(\ref{derivative})$ by $\alpha_i$ and then summing from 1 to $k^*$ yields
\begin{eqnarray*}
2 n^{-4/d}(k^*)^{1+2/d}\sum_{i=1}^{k^*}\alpha_i w_{ni} + 2\lambda + \nu k^* &=& 0\\
2 n^{-4/d}\sum_{i=1}^{k^*}\alpha_i w_{ni} \sum_{i=1}^{k^*} \alpha_i^2 + 2\lambda \sum_{i=1}^{k^*}\alpha_i w_{ni} + \nu (k^*)^{1+2/d} &=& 0.
\end{eqnarray*}
Therefore, we have
\begin{equation}
w_{ni}^{*} = \frac{1}{k^*} + \frac{(k^*)^{4/d}-(k^*)^{2/d}\alpha_i}{\sum_{i=1}^{k^*} \alpha_i^2+\lambda n^{4/d} -(k^*)^{1+4/d}} \label{optweight0}
\end{equation}
Here $w_{ni}^{*}$ is decreasing in $i$ since $\alpha_i$ is increasing in $i$ and $\sum_{i=1}^{k^{*}}\alpha_i^2>(k^{*})^{1+4/d}$ from Lemma \ref{alpha}. Next we solve for $k^*$. According to the definition of $k^{*}$, we only need to find $k$ such that $w_{nk}^{*}=0$. Using the results from Lemma \ref{alpha}, solving this equation reduces to solving $k^*$ such that
$$
(1+\frac{2}{d})(k^*-1)^{2/d} \le \lambda n^{4/d}(k^*)^{-1-2/d} + \frac{(d+2)^2}{d(d+4)}(k^*)^{2/d}\{1+O(\frac{1}{k^*})\} \le (1+\frac{2}{d})(k^*)^{2/d}.
$$
Therefore, for large $n$, we have
\begin{equation}
k^*= \Big\lfloor \Big\{\frac{d(d+4)}{2(d+2)}\Big\}^{\frac{d}{d+4}}\lambda^{\frac{d}{d+4}}n^{\frac{4}{d+4}}\Big\rfloor.\nonumber
\end{equation}
Plugging $k^*$ and the result $(\ref{sum})$ in Supplementary into $(\ref{optweight0})$ yields the optimal weight. \hfill $\blacksquare$
\subsection{Proof of Theorem \ref{thm:upperCIS}}
\label{sec:proofuppercis}
Following the proofs of Lemma 3.1 in \citet{AT07}, we consider the sets $A_j \subset {\cal R}$
\begin{eqnarray*}
A_0 &=& \{x\in {\cal R}: 0<|\eta(x)-1/2|\le \delta\},\\
A_j &=& \{x\in {\cal R}: 2^{j-1}\delta < |\eta(x)-1/2|\le 2^j\delta\} \textrm{~for~} j\ge 1.
\end{eqnarray*}
For the classification procedure $\Psi(\cdot)$, we have
\begin{eqnarray*}
\textrm{CIS}(\Psi) = {\mathbb E}[\ind{\widehat{\phi}_{n1}(X) \ne \widehat{\phi}_{n2}(X)}],
\end{eqnarray*}
where $\widehat{\phi}_{n1}$ and $\widehat{\phi}_{n2}$ are classifiers obtained by applying $\Psi(\cdot)$ to two independently and identically distributed samples ${\cal D}_1$ and ${\cal D}_2$, respectively. Denote the Bayes classifier $\phi^{\textrm{Bayes}}$, we have
\begin{eqnarray*}
\textrm{CIS}(\Psi) &=& 2 {\mathbb E}[\ind{\widehat{\phi}_{n1}(X) = \phi^{\textrm{Bayes}}(X), \widehat{\phi}_{n2}(X)\ne \phi^{\textrm{Bayes}}(X)}]\\
&=& 2 {\mathbb E}[ \{1 - \ind{\widehat{\phi}_{n1}(X) \ne \phi^{\textrm{Bayes}}(X)}\} \ind{\widehat{\phi}_{n2}(X)\ne \phi^{\textrm{Bayes}}(X)}] \\
&=& 2 {\mathbb E}_X[ {\mathbb P}_{{\cal D}_1}(\widehat{\phi}_{n1}(X) \ne \phi^{\textrm{Bayes}}(X)|X) - \{{\mathbb P}_{{\cal D}_1}(\widehat{\phi}_{n1}(X) \ne \phi^{\textrm{Bayes}}(X)|X)\}^2] \\
&\le & 2 {\mathbb E} [ \ind{\widehat{\phi}_{n1}(X) \ne \phi^{\textrm{Bayes}}(X)} ],
\end{eqnarray*}
where the last equality is due to the fact that ${\cal D}_1$ and ${\cal D}_2$ are independently and identically distributed. For ease of notation, we will denote $\widehat{\phi}_{n1}$ as $\widehat{\phi}_{n}$ from now on. We further have
\begin{eqnarray*}
\textrm{CIS}(\Psi) &\le & 2 \sum_{j=0}^\infty {\mathbb E} [ \ind{\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)} \ind{X\in A_j}]\\
&\le& 2 {\mathbb P}_X(0<|\eta(X)-1/2|\le \delta) + 2 \sum_{j\ge 1} {\mathbb E} [ \ind{\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)} \ind{X\in A_j}].
\end{eqnarray*}
Given the event $\{\widehat{\phi}_{n} \ne \phi^{\textrm{Bayes}}\}\cap\{|\eta-1/2|>2^{j-1}\delta\}$, we have $|\widehat{\eta}_n - \eta| \ge 2^{j-1}\delta$. Therefore, for any $j\ge 1$, we have
\begin{eqnarray*}
&& {\mathbb E} [ \ind{\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)} \ind{X\in A_j}]\\
&\le & {\mathbb E} [ \ind{|\widehat{\eta}_n(X) - \eta(X)| \ge 2^{j-1}\delta} \ind{2^{j-1}\delta < |\eta(X)-1/2|\le 2^j\delta}]\\
&\le & {\mathbb E}_X [ {\mathbb P}_{\cal D}(|\widehat{\eta}_n(X) - \eta(X)| \ge 2^{j-1}\delta|X) \ind{0 < |\eta(X)-1/2|\le 2^j\delta}]\\
&\le & C_1 \exp(-C_2 a_n (2^{j-1}\delta)^2) {\mathbb P}_X (0 < |\eta(X)-1/2|\le 2^j\delta)\\
&\le & C_1 \exp(-C_2 a_n (2^{j-1}\delta)^2) C_0 (2^j\delta)^{\alpha},
\end{eqnarray*}
where the last inequality is due to margin assumption $(\ref{margin})$ and condition $(\ref{exponential})$.
Taking $\delta=a_n^{-1/2}$, we have
$$
\textrm{CIS}(\Psi) \le C_0 a_n^{-\alpha/2} + C_0C_1 a_n^{-\alpha/2} \sum_{j\ge 1} 2^{\alpha j +1} e^{-C_2 4^{j-1}} \le C a_n^{-\alpha/2},
$$
for some $C>0$ depending only on $\alpha,C_0,C_1$ and $C_2$. \hfill $\blacksquare$\\
\subsection{Proof of Theorem \ref{thm:lowerCIS}} According to the proof of Theorem \ref{thm:upperCIS}, we have
\begin{eqnarray*}
\textrm{CIS}(\Psi) = 2 \left\{ \mathbb E_X[ \mathbb P_{\cal D}(\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)|X)] - \mathbb E_X[\{\mathbb P_{\cal D}(\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)|X)\}^2]\right\}.
\end{eqnarray*}
\citet{AT07} showed that when $\alpha\gamma\le d$, the set of probability distribution ${\cal P}_{\alpha,\gamma}$ contains a $(m,w,b,b')$-hypercube with $w=C_3 q^{-d}$, $m=\lfloor C_4 q^{d-\alpha\gamma}\rfloor$, $b=b'=C_5q^{-\gamma}$ and $q=\lfloor C_6n^{1/(2\gamma+d)} \rfloor$, with some constants $C_i\ge 0$ for $i=3,\ldots,6$ and $C_6\le 1$. Therefore, Lemma \ref{lemma:assouad} implies that the first part is bound, that is,
\begin{eqnarray*}
&& \sup_{ P\in {\cal P}_{\alpha,\gamma}} \mathbb E_X[ \mathbb P_{\cal D}(\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)|X)]\\
&=& \sup_{P\in {\cal P}_{\alpha,\gamma}} \mathbb E_{\cal D} [\mathbb P_X(\widehat{\phi}_n(X)\ne \phi^{\textrm{Bayes}}(X))]\\
&\ge &\frac{mw}{2} [1-b\sqrt{nw}]\\
&=& (1-C_6)C_3C_4C_5 n^{-\alpha\gamma/(2\gamma+d)}.
\end{eqnarray*}
To bound the second part, we again consider the sets $A_j$ defined in Appendix \ref{sec:proofuppercis}. On the event $\{\widehat{\phi}_{n} \ne \phi^{\textrm{Bayes}}\}\cap\{|\eta-1/2|>2^{j-1}\delta\}$, we have $|\widehat{\eta}_n - \eta| \ge 2^{j-1}\delta$. Letting $\delta = a_n^{-1/2}$ leads to
\begin{eqnarray*}
&& \mathbb E_X[\{\mathbb P_{\cal D}(\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)|X)\}^2]\\
&= & \sum_{j=0}^{\infty} \mathbb E_X [ \{\mathbb P_{\cal D}(\{\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)\}|X)\}^2 \ind{X\in A_j}]\\
&\le & \mathbb P_X (0 < |\eta(X)-1/2|\le \delta) + \sum_{j=1}^{\infty} \mathbb E_X [ \{\mathbb P_{\cal D}(\{\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)\}|X)\}^2 \ind{X\in A_j}]\\
&\le & \mathbb P_X (0 < |\eta(X)-1/2|\le \delta) + \sum_{j\ge 1} C_1e^{-2C_2 4^{j-1}} \mathbb P_X (0 < |\eta(x)-1/2|\le 2^j\delta)\\
&\le & C_0 a_n^{-\alpha/2} + C_0C_1 a_n^{-\alpha/2} \sum_{j\ge 1} 2^{\alpha j} e^{-2C_2 4^{j-1}}\\
&\le & C_7 a_n^{-\alpha/2},
\end{eqnarray*}
for some positive constant $C_7$ depending only on $\alpha, C_0, C_1, C_2$. When $a_n=n^{2\gamma/(2\gamma+d)}$, we have
$$
\mathbb E_X [ (\mathbb P_{\cal D}(\widehat{\phi}_{n}(X) \ne \phi^{\textrm{Bayes}}(X)|X))^2 ] \le C_7 n^{-\alpha\gamma/(2\gamma+d)}.
$$
By properly choosing constants $C_i$ such that $(1-C_6)C_3C_4C_5 - C_7>0$, we have
$$
\textrm{CIS}(\Psi) \ge 2 [(1-C_6)C_3C_4C_5 - C_7] n^{-\alpha\gamma/(2\gamma+d)} \ge C' n^{-\alpha\gamma/(2\gamma+d)},
$$
for a constant $C'>0$. This concludes the proof of Theorem \ref{thm:lowerCIS}. \hfill $\blacksquare$
\subsection{Proof of Theorem \ref{thm:upperCISsnn}}
\label{sec:proofupperCISsnn}
According to our Theorem \ref{thm:upperCIS} and the proof of Theorem 1 in the supplementary of \citet{S12}, it is sufficient to show that for any $\alpha\ge 0$ and $\gamma\in (0,2]$, there exist positive constants $C_1,C_2$ such that for all $\delta>0$, $n\ge 1$ and $\bar{P}$-almost all $x$,
\begin{equation}
\sup_{P\in {\cal P}_{\alpha,\gamma}} \mathbb P_{\cal D}\Big(|S_n^*(x)-\eta(x)| \ge \delta\Big) \le C_1 \exp(-C_2n^{2\gamma/(2\gamma+d)}\delta^2). \label{ntp:exponential}
\end{equation}
where $S_n^*(x)=\sum_{i=1}^n w_{ni}^*\ind{Y_{(i)}=1}$ with the optimal weight $w_{ni}^*$ defined in Theorem \ref{thm:optimal} and $k^*\asymp n^{2\gamma/(2\gamma+d)}$.
According to Lemma \ref{alpha}, we have
\begin{eqnarray*}
\sum_{i=1}^{k^*}(w_{ni}^*)^2 = \frac{2(d+2)}{(d+4)k^*}\{1+O((k^*)^{-1})\} \le C_8 n^{-2\gamma/(2\gamma+d)},
\end{eqnarray*}
for some constant $C_8>0$.
Denote $\mu^*_n(x)={\mathbb E}\{S^*_n(x)\}$. According to the proof of Theorem 1 in the supplement of \citet{S12}, there exist $C_9,C_{10}>0$ such that for all $P\in {\cal P}_{\alpha,\gamma}$ and $x\in {\cal R}$,
\begin{eqnarray}
|\mu^*_n(x) - \eta(x)| &\le& \left|\sum_{i=1}^n w_{ni}^* \mathbb E\{\eta(X_{(i)}) - \eta_x(X_{(i)})\}\right|+\left|\sum_{i=1}^n w_{ni}^* \mathbb E\{\eta_x(X_{(i)})\} - \eta(x)\right| \nonumber\\
&\le& L\sum_{i=1}^n w_{ni}^* \mathbb E\{\|X_{(i)}-x\|^{\gamma}\} + \left|\sum_{i=1}^n w_{ni}^* \mathbb E\{\eta_x(X_{(i)})\} - \eta(x)\right|\nonumber\\
&\le & C_9 \sum_{i=1}^n w_{ni}^*\Big(\frac{i}{n}\Big)^{\gamma/d}\nonumber\\
&\le & C_{10} n^{-\gamma/(2\gamma+d)}. \label{halfdelta}
\end{eqnarray}
The Hoeffding's inequality says that if $Z_1,\ldots,Z_n$ are independent and $Z_i\in [a_i,b_i]$ almost surely, then we have
$$
\mathbb P\left(\Big|\sum_{i=1}^n Z_i-\mathbb E\Big[\sum_{i=1}^n Z_i\Big]\Big|\ge t\right) \le 2\exp\Big(-\frac{2t^2}{\sum_{i=1}^n(b_i-a_i)^2}\Big).
$$
Let $Z_i=w^*_{ni}\ind{Y_{(i)}=1}$ with $a_i=0$ and $b_i=w^*_{ni}$. According to $(\ref{halfdelta})$, we have that for $\delta\ge 2C_{10} n^{-\gamma/(2\gamma+d)}$ and for $\bar{P}$-almost all $x$,
\begin{eqnarray*}
\sup_{P\in {\cal P}_{\alpha,\gamma}} \mathbb P_{\cal D}\Big(|S_n^*(x)-\eta(x)| \ge \delta\Big) &\le& \sup_{P\in {\cal P}_{\alpha,\gamma}} \mathbb P_{\cal D}\Big(|S_n^*(x)-\mu^*_n(x)| \ge \delta/2 \Big)\\
&\le & 2 \exp \{-n^{2\gamma/(2\gamma+d)}\delta^2/(2C_8)\},
\end{eqnarray*}
which implies $(\ref{ntp:exponential})$ directly. \hfill $\blacksquare$
|
\section{Introduction}
Cavity optomechanics, exploring the interaction between light fields and
mechanical motions, has attracted a lot of attention in the past few years for
its potential application in the ultrasensitive detection of tiny mass, force,
and displacement \cite{A1,A2,A3,A4}. One standard and simplest optomechanical
setup is a Fabry-Perot cavity with one end mirror being a micro- or
nano-mechanical vibrating object \cite{B1,B2,B3}. Other various optomechanical
experimental system are designed and investigated such as silica toroidal
optical microresonators \cite{C1,C2,C3}, photonic crystal cavities \cite{D1,D2},
micromechanical membranes \cite{E1,E2}, typical optomechanical cavities confining cold
atoms \cite{f1,f2}, superconducting circuits \cite{g1,g2}, and so on.
Typically, when driving an otomechanical cavity by a red-detuned laser, the
mechanical oscillator can be cooled to its quantum
ground-state \cite{h1,h2,h3}. Moreover,
in this red-detuned regime, some well-known phenomena in atomic ensemble can find their analogy in optomechanical system. Specifically, under a strong driving, normal mode splitting \cite{j1,j2}(called Autler-Townes effects in atomic physics) can be observed. On the contrary, for a relatively weak driving (much less than the cavity dissipation rate), an electromagnetically induced transparency like phenomenon, called optomechanically induced transparency\cite{L1,L2,L3}, has been theoretically predict and experimentally verified. This phenomenon can be used to slow down and even stop light signals
\cite{M1,M2} in the long-lived mechanical vibrations.
On the other hand,
when a driving laser applied on the mechanical blue sideband, the
mechanical element of an optomecanical system can be heated, leading to phonon lasing \cite{i1,i2,i3}and probe amplification \cite{p1,p2,N1,N2}.
In our previous work, we have investigated coherent perfect transmission and absorption in a double-cavity optomechanical system driven by two pump fields on red mechanical sideband \cite{q1}. While in this paper, we study the
optomechanically induced amplification and perfect transparency in the same system driven under a different type of drive. We find that if driving the double-cavity
optomechanical system by a red sideband laser from one side and a blue sideband one from the other side and appropriately manipulating the amplitudes of them, optomechanically induced amplification phenomenon can occur for a nearly resonant weak signal
field (probe field). In addition, by adjusting the control fields, an interesting perfect
optomechanically induced transparency (with transmission coefficient rigorously equal to 1) can be realized under the same type of drive. When this perfect transmission occur, quantum coherence process due to the double-driving can totally suppress the decoherence due to the dissipation of mechanical resonator. This double-driving device could be used to
realize signal quantum amplifier, quantum switch, quantum memory and so on.
The rest of this paper is organized as follows. In Section II, we introduce
the double-cavity optomechanical model, obtain the equations of motion for the
mechanical resonator and the two cavity modes, and solve them and obtain the
output fields. In Section III, we show how to realize perfect optomechanically
induced transparency even though with big mechanical decay rate $\gamma_{m}$.
In Section IV, we show how to realize optomechanically induced amplification about the
weak signal field (probe field), meanwhile, the system holds below the phonon
lasing threshold. And the conclusions are given in the Section V.
\section{Model and Equations}
\begin{figure}[ptbh]
\includegraphics[width=0.45\textwidth]{Fig1.eps}\caption{(Color online) A
double-cavity optomechanical system with a mechanical resonator (MR) inserted
between two fixed mirrors. The two cavities have identical cavity lengths $L$
and mode frequencies $\omega_{0}$ in the absence of radiation pressure.
Coupling field and driving field with frequencies $\omega_{c}, \omega_{d}$ and
amplitudes $\varepsilon_{c}, \varepsilon_{d}$ respectively, act upon opposite
sides of the double-cavity system. The probe field with frequency $\omega_{p}$
and amplitude $\varepsilon_{p}$ is injected into the left optical cavity.
\label{Fig1
\end{figure}
We consider a double-cavity hybrid system with one mechanical resonator (MR)
of perfect reflection inserted between two fixed mirrors of partial
transmission (see Fig. 1). The MR has an eigen frequency $\omega_{m}$ and a
decay rate $\gamma_{m}$ and thus exhibits a mechanical quality factor
$Q=\omega_{m}/\gamma_{m}$. Two identical optical cavities of lengths $L$ and
frequencies $\omega_{0}$ are got when the MR is at its equilibrium position in
the absence of external excitation. We describe the two optical modes,
respectively, by annihilation (creation) operators $c_{1}$ ($c_{1}^{\dagger}$)
and $c_{2}$ ($c_{2}^{\dagger}$) while the only mechanical mode by $b$
($b^{\dagger}$). These annihilation and creation operators are restricted by
the commutation relation $[c_{i},c_{i}^{\dagger}]=1$ ($i=1,2$) , $[c_{1
,c_{2}]=0$, and $[b,b^{\dagger}]=1$. Two coupling fields are used to drive the
double-cavity system from either left or right fixed mirrors with their
amplitudes denoted by $\varepsilon_{c}=\sqrt{2\kappa\wp_{c}/(\hbar\omega_{c
)}$ and $\varepsilon_{d}=\sqrt{2\kappa\wp_{d}/(\hbar\omega_{d})}$ and one
probe field is injected into the left optical cavity with amplitude denoted by
$\varepsilon_{p}=\sqrt{2\kappa\wp_{p}/(\hbar\omega_{p})}$. Here $\wp_{c}$,
$\wp_{d}$ and $\wp_{p}$, are relevant field powers,\ $\kappa$ is the common
decay rate of both cavity modes, and $\omega_{c}$, $\omega_{d}$, and
$\omega_{p}$, are relevant field frequencies. Then the total Hamiltonian in
the rotating-wave frame of frequency $\omega_{c}+\omega_{d}$ can be written
as
\begin{align}
H & =\hbar\Delta_{c}c_{1}^{\dagger}c_{1}+\hbar\Delta_{d}c_{2}^{\dagger
c_{2}+\hbar g_{0}(c_{2}^{\dagger}c_{2}-c_{1}^{\dagger}c_{1})(b^{\dagger
}+b)\label{Eq1}\\
& +\hbar\omega_{m}b^{\dagger}b+i\hbar\varepsilon_{c}(c_{1}^{\dagger
-c_{1})+i\hbar\varepsilon_{d}(c_{2}^{\dagger}-c_{2})\nonumber\\
& +i\hbar(c_{1}^{\dagger}\varepsilon_{p}e^{-i\delta t}-c_{1}\varepsilon
_{p}^{\ast}e^{i\delta t})\nonumber
\end{align}
with $\Delta_{c}=\omega_{0}-\omega_{c}$ ($\Delta_{d}=\omega_{0}-\omega_{d}$)
being the detuning between cavity modes and coupling field (driving field),
$\delta=\omega_{p}-\omega_{c}$ being the detuning between the probe field and
the coupling field, and $g_{0}=\frac{\omega_{0}}{L}\sqrt{\frac{\hbar
{2m\omega_{m}}}$ being the hybrid coupling constant between mechanical and
optical modes.
The dynamics of the system is described by the quantum Langevin equations
for relevant annihilation operators of mechanical and optical mode
\begin{align}
\dot{b} & =-i\omega_{m}b-ig_{0}(c_{2}^{\dagger}c_{2}-c_{1}^{\dagger
c_{1})-\frac{\gamma_{m}}{2}b+\sqrt{\gamma_{m}}b_{in},\label{Eq2}\\
\dot{c}_{1} & =-[\kappa+i\Delta_{c}-ig_{0}(b^{\dagger}+b)]c_{1
+\varepsilon_{c}+\varepsilon_{p}e^{-i\delta t}+\sqrt{2\kappa}c_{1
^{in},\nonumber\\
\dot{c}_{2} & =-[\kappa+i\Delta_{d}+ig_{0}(b^{\dagger}+b)]c_{2
+\varepsilon_{d}+\sqrt{2\kappa}c_{2}^{in}\nonumber
\end{align}
with $b_{in}$ being the thermal noise on the MR with zero mean value,
$c_{1}^{in}$ ($c_{2}^{in}$) is the input quantum vacuum noise from the left
(right) cavity with zero mean value. Because we deal with the mean response of
the system, we do not include these noise terms in the discussion that
follows. In the absence of probe field $\varepsilon_{p}$, Eq.s (2) can be
solved with the factorization assumption $\left\langle bc_{i}\right\rangle
=\left\langle b\right\rangle \left\langle c_{i}\right\rangle $ to generate the
steady-state mean value
\begin{align}
\langle b\rangle & =b_{s}=\frac{-ig_{0}(\left\vert c_{2s}\right\vert
^{2}-\left\vert c_{1s}\right\vert ^{2})}{\frac{\gamma_{m}}{2}+i\omega_{m
},\label{Eq3}\\
\langle c_{1}\rangle & =c_{1s}=\frac{\varepsilon_{c}}{\kappa+i\Delta_{1
},\nonumber\\
\langle c_{2}\rangle & =c_{2s}=\frac{\varepsilon_{d}}{\kappa+i\Delta_{2
}\nonumber
\end{align}
with $\Delta_{1,2}=\Delta_{c,d}\mp g_{0}(b_{s}+b_{s}^{\ast})$ denoting the
effective detunings between cavity modes and coupling field, driving field
when the membrane oscillator deviates from its equilibrium position. Note in
particular, that $g_{0}\left\vert b_{s}\right\vert $ is typically very small
as compared to $\omega_{m}$ and becomes even exactly zero in the case of
$\left\vert c_{1s}\right\vert =\left\vert c_{2s}\right\vert $ ($\left\vert
\varepsilon_{c}\right\vert =\left\vert \varepsilon_{d}\right\vert $).
In the presence of probe field, however, we can write each operator as the sum
of its mean value and its small fluctuation ($b=b_{s}+\delta b,c_{1
=c_{1s}+\delta c_{1},c_{2}=c_{2s}+\delta c_{2}$) to solve Eq. (2) when the
coupling field and the driving field are sufficiently strong. Then keeping
only the linear terms of fluctuation operators and moving into an interaction
picture by introducing $\delta b\rightarrow\delta be^{-i\omega_{m}t}$, $\delta
c_{1}\rightarrow\delta c_{1}e^{-i\Delta_{1}t}$, $\delta c_{2}\rightarrow\delta
c_{2}e^{-i\Delta_{2}t}$, we obtain the linearized quantum Langevin equation
\begin{align}
\delta\dot{b} & =-ig_{0}(c_{2s}^{\ast}\delta c_{2}e^{-i(\Delta_{2}-\omega
_{m})t}-c_{1s}^{\ast}\delta c_{1}e^{-i(\Delta_{1}-\omega_{m})t})\label{Eq4}\\
& -ig_{0}(c_{2s}\delta c_{2}^{\dagger}e^{i(\Delta_{2}+\omega_{m})t
-c_{1s}\delta c_{1}^{\dagger}e^{i(\Delta_{1}+\omega_{m})t})-\frac{\gamma_{m
}{2}\delta b,\nonumber\\
\delta\dot{c}_{1} & =-\kappa\delta c_{1}+ig_{0}c_{1s}(\delta be^{-i(\omega
_{m}-\Delta_{1})t}+\delta b^{\dagger}e^{i(\omega_{m}+\Delta_{1})t})\nonumber\\
& +\varepsilon_{p}e^{-i(\delta-\Delta_{1})t},\nonumber\\
\delta\dot{c}_{2} & =-\kappa\delta c_{2}-ig_{0}c_{2s}(\delta be^{-i(\omega
_{m}-\Delta_{2})t}+\delta b^{\dagger}e^{i(\omega_{m}+\Delta_{2})t}).\nonumber
\end{align}
If the coupling field drives at the mechanical red sideband while the driving
field drives at the blue sideband ($\Delta_{1}\approx\omega_{m}$, $\Delta
_{2}\approx-\omega_{m}$), the hybrid system is operating in the resolved
sideband regime ($\omega_{m}>>\kappa$), the membrane oscillator has a high
mechanical quality factor ($\omega_{m}>>\gamma_{m}$), and the mechanical
frequency $\omega_{m}$ is much larger than $g_{0}\left\vert c_{1s}\right\vert
$ and $g_{0}\left\vert c_{2s}\right\vert $, Eq.s (4) will be simplified to
\begin{align}
\delta\dot{b} & =-ig_{0}(c_{2s}\delta c_{2}^{\dagger}-c_{1s}^{\ast}\delta
c_{1})-\frac{\gamma_{m}}{2}\delta b,\label{Eq5}\\
\delta\dot{c}_{1} & =-\kappa\delta c_{1}+ig_{0}c_{1s}\delta b+\varepsilon
_{L}e^{-ixt},\nonumber\\
\delta\dot{c}_{2} & =-\kappa\delta c_{2}-ig_{0}c_{2s}\delta b^{\dagger
}\nonumber
\end{align}
with $x=\delta-\omega_{m}$. We can examine the expectation values of small
fluctuations by the following three coupled dynamic equations
\begin{align}
\left\langle \delta\dot{b}\right\rangle & =-ig_{0}(c_{2s}\left\langle \delta
c_{2}^{\dagger}\right\rangle -c_{1s}^{\ast}\left\langle \delta c_{1
\right\rangle )-\frac{\gamma_{m}}{2}\left\langle \delta b\right\rangle
,\label{Eq6}\\
\left\langle \delta\dot{c}_{1}\right\rangle & =-\kappa\left\langle \delta
c_{1}\right\rangle +ig_{0}c_{1s}\left\langle \delta b\right\rangle
+\varepsilon_{p}e^{-ixt},\nonumber\\
\left\langle \delta\dot{c}_{2}\right\rangle & =-\kappa\left\langle \delta
c_{2}\right\rangle -ig_{0}c_{2s}\left\langle \delta b^{\dagger}\right\rangle.
\nonumber
\end{align}
We assume the steady-state solutions of above equations have form: $\langle\delta s\rangle=\delta s_{+}e^{-ixt}+\delta s_{-}e^{ixt}$
with $s=b,c_{1},c_{2}$. Then it is straightforward to obtain the following
results
\begin{align}
\delta b_{+} & =\frac{iG\varepsilon_{p}}{(\kappa-ix)(\frac{\gamma_{m}
{2}-ix)+G^{2}(1-n^{2})},\\
\delta c_{1+} & =\frac{\varepsilon_{p}[-n^{2}G^{2}+(\kappa-ix)(\frac
{\gamma_{m}}{2}-ix)]}{(\kappa-ix)^{2}(\frac{\gamma_{m}}{2}-ix)+G^{2
(1-n^{2})(\kappa-ix)},\nonumber\\
\delta c_{2-} & =\frac{-nG^{2}\varepsilon_{p}}{(\kappa+ix)^{2}(\frac
{\gamma_{m}}{2}+ix)+G^{2}(1-n^{2})(\kappa+ix)},\nonumber
\end{align}
where $G=g_{0}c_{1s}$ is the effective optomechanical coupling
rate and $\left\vert c_{2s}/c_{1s}\right\vert ^{2}=n^{2}$ is the photon number
ratio of two cavity modes. In deriving Eqs. (7), we have also assumed that
$c_{1s,2s}$ is real-valued without loss of generality.
Based on Eqs. (7), we can further determine the left-hand output field
$\varepsilon_{outL}$ and the right-hand output field $\varepsilon_{outR}$
through the following input-output relation \cite{Walls} \bigski
\begin{align}
\varepsilon_{outL} & =2\kappa\langle\delta c_{1}\rangle-\varepsilon
_{p}e^{-ixt}\label{Eq8}\\
\varepsilon_{outR} & =2\kappa\langle\delta c_{2}\rangle,\nonumber
\end{align}
where the oscillating terms can be removed if we set $\varepsilon
_{outL}=\varepsilon_{outL+}e^{-ixt}+\varepsilon_{outL-}e^{ixt}$ and
$\varepsilon_{outR}=\varepsilon_{outR+}e^{-ixt}+\varepsilon_{outR-}e^{ixt}$.
Note that the output components $\varepsilon_{outL+}$ and $\varepsilon
_{outR-}$ have the same frequency $\omega_{p}$\ as the input probe
fields $\varepsilon_{p}$, while the output components $\varepsilon_{outL-}$ and
$\varepsilon_{outR+}$ are generated at frequencies $2\omega_{c}-\omega_{p}$ and $2\omega_{d}-\omega_{p}$, respectively, in a nonlinear wave-mixing process of optomechanical
interaction. Then with Eqs. (8) we can obtain
\begin{align}
\varepsilon_{outL+} & =2\kappa\delta c_{1+}-\varepsilon_{p},\label{Eq9}\\
\varepsilon_{outR-} & =2\kappa\delta c_{2-}\nonumber
\end{align}
oscillating at frequency $\omega_{p}$ of our special interest.
In this paper, we discuss the perfect optomechanically induced amplification and transparency
under the realistic parameters in a
optomechanical experiment \cite{j2}. That is, $L=25$ mm, $m=145$
ng, $\kappa=2\pi\times215$ kHz, $\omega_{m}=2\pi\times947$ kHz, and
$\gamma_{m}=2\pi\times141$ Hz. In addition, the laser wavelength is $\lambda$
$=$ $2\pi c/\omega_{c}=1064$ nm and the mechanical quality factor is $Q=$
$\omega_{m}/\gamma_{m}=6700$.
\section{Perfect optomechanically induced transparency}
Now we study the perfect optomechanically induced transparency for the probe
field. The quadrature of the optical components with
frequency $\omega_{p}$ in the output field can be defined as $\varepsilon_{T}=2\kappa\delta c_{1+}/\varepsilon_{p}$
\cite{L2} . Specifically, it can be written as
\begin{align}
\varepsilon_{T} =\frac{2\kappa[-n^{2}G^{2}+(\kappa-ix)(\frac{\gamma_{m}
{2}-ix)]}{(\kappa-ix)^{2}(\frac{\gamma_{m}}{2}-ix)+G^{2}(1-n^{2})(\kappa-ix)},
\end{align}
whose real and imaginary part ($Re[\varepsilon_{T}]$ and $Im[\varepsilon_{T}]$)
represent the absorptive and dispersive behavior of the optomechanical system,
respectively. It is well-known that in a standard optomechanical system with single optical cavity, the optomechanically induced transparency
dip is not perfect as the decay $\gamma_{m}$ of the mechanical resonator is
not zero. However, we can see from Eq.(10) that, in the double-cavity optomechanical system studied here, if setting the ratio $n=\sqrt{\gamma_{m
\kappa/2G^{2}}$, the optomechanically induced transparency dip will be perfect
even though remarkale mechanical decay $\gamma_{m}$ exists.
\begin{figure}[ptbh]
\centering \includegraphics[width=0.45\textwidth]{Fig2.eps}\caption{The real
part of $\varepsilon_{T}$ vs. the normalized frequency detuning $x/\kappa$:
$n=0$ (red-dashed) and $n=0.7$ (black-solid) with $\gamma_{m}=2\pi\times14.1$
kHz and $\wp_{c}=1mW$. In the inset: $n=0.7$, $\gamma_{m}=2\pi\times141$ Hz
and $\wp_{c}=1mW$.
\label{Fig2
\end{figure}
\begin{figure}[ptbh]
\centering \includegraphics[width=0.45\textwidth]{Fig3.eps}\caption{The imaginary
part of $\varepsilon_{T}$ vs. the normalized frequency detuning $x/\kappa$:
$n=0$ (red-dashed) and $n=0.7$ (black-solid) with $\gamma_{m}=2\pi\times14.1$
kHz and $\wp_{c}=1mW$.
\label{Fig3
\end{figure}
To see this clearly, in Fig. 2, we plot the $Re[\varepsilon_{T}]$ versus the
normalized frequency $x/\kappa$ with $\gamma_{m}=2\pi\times14.1$ kHz and
$\wp_{c}=1mW$ for different $n$. We can see form Fig. 2 that when $n=0$ (i.e. the usual optomechanically induced transparency case), the
optomechanically induced transparency dip will become shallow with a large
mechanical decay $\gamma_{m}$ (red-dashed). However, when an additional blue-sideband driving field satisfying the condition $n=\sqrt{\gamma_{m}\kappa/2G^{2}}\approx0.7$ applied, the transparency dip will become
perfect, exhibiting totally transmission of probe laser (black-solid). Physically,
it means that the dissipative energy through the decay
$\gamma_{m}$ of the mechanical resonator can be compensated by applying the
right-hand driving field with amplitude $\varepsilon_{d}=\varepsilon_{c
\sqrt{\gamma_{m}\kappa/2G^{2}}$ and the blue mechanical sideband frequency.
When $\omega_{p}\approx\omega_{0}$, $n=\sqrt{\gamma_{m}\kappa/2G^{2}}$ and the
beat frequency $\omega_{p}-\omega_{c}=\omega_{m} (x=0)$, thus, the MR is
driven by a force oscillating at its eigenfrequency $\omega_{m}$ and the
resonator starts to oscillate coherently. This motion will generate photons
with frequency $\omega_{p}$ that interfere destructively with the probe beam,
leading to a optomechanically induced transparency dip.
In Fig. 3, we plot the the dispersion curve $Im[\varepsilon_{T}]$ versus the
normalized frequency $x/\kappa$ with $\gamma_{m}=2\pi\times14.1$ kHz and
$\wp_{c}=1mW$ for different $n$. Clearly, the curve with $n=0.7$ (black-solid) is much steeper than the one with $n=0$ (red-dashed) in the vicinity of $x=0$. It means that we can easily control the dispersive behavior of the optomechanical system by applying the blue-detuned driving field with amplitude $\varepsilon_{d}=n\varepsilon_{c}$, which can possibly be used to control slow light in optomechanical systems.
\section{optomechanically induced amplification}
In this section, we study the optomechanically induced amplification in this double-cavity optomechanical system. If the ratio $n>\sqrt{\gamma_{m}\kappa/2G^{2}}$, we find the $Re[\varepsilon
_{T}]$ will become negative in the vicinity of $x=0$ (see the inset in Fig.
2). It means that optomechanically induced gain (amplification) can be
realized in this double-cavity system by applying a blue-detuned driving field to the right-side cavity
with amplitude $\varepsilon_{d}=n\varepsilon_{c}$. Note that when the system works under the condition
$x=0$, $n=\sqrt
{1+\frac{\gamma_{m}\kappa}{2G^{2}}}$, $Re[\varepsilon
_{T}]$ will be divergent. In addition, the system will work into the
parametric instability regime as $n\gtrsim1$ when the input power $\wp_{c
=1$mW, and so we limit ourselves to the case where $n\leq1$.
\begin{figure}[ptbh]
\centering \includegraphics[width=0.45\textwidth]{Fig4.eps}\caption{The
normalized mechanical oscillation $|\kappa\delta b_{+}/\varepsilon_{p}|^{2}$
vs. the normalized frequency detuning $x/\kappa$: $n=0$ (black-dotted),
$n=0.7$ (green-dotted-dashed), $n=0.8$ (blue-dashed), and $n=0.9$ (red-solid)
with $\wp_{cL}=1mW$. In the inset, we plot the normalized mechanical
oscillation $|\kappa\delta b_{+}/\varepsilon_{p}|^{2}$ vs. the ratio $n$.
\label{Fig4
\end{figure}
In Fig. 4, we plot the mechanical oscillation $|\kappa\delta b_{+
/\varepsilon_{p}|^{2}$ (normalized to probe field $\varepsilon_{p}$) versus
the normalized frequency $x/\kappa$ for different $n$. In the inset we plot
the $|\kappa\delta b_{+}/\varepsilon_{p}|^{2}$ as a function of $n$ for $x=0$.
It can be seen clearly from Fig. 4 that the mechanical oscillation peak value
locates at $x=0$, and increases with $n$ [$n=0$
(black-dotted), $n=0.7$ (green-dotted-dashed), $n=0.8$ (blue-dashed), $n=0.9$
(red-solid)]. And when $n$ increases up to 1, the mechanical oscillation peak
value will increase approximately to $6.1\times10^{5}$ (see the inset in Fig.
4). It means that the optomechanical effect will become stronger for bigger
$n$ (less than or equal to 1) when $\omega_{p}-\omega_{c}=\omega_{m} (x=0)$
and $\omega_{d}-\omega_{0}=\omega_{m}$. The reason for this is that, the
blue-mechanical sideband (heating sideband) of right-hand cavity generating
much phonons which will be absorbed by the Anti-Stokes processes in left-hand
cavity for the red-mechanical sideband (cooling sideband). Then, the
optomechanical effect of the double-cavity system is resonantly enhanced.
\begin{figure}[ptbh]
\centering \includegraphics[width=0.45\textwidth]{Fig5.eps}\caption{The
normalized left-hand output energy $|\varepsilon_{outL+}/\varepsilon_{p}|^{2}$
vs. the normalized frequency detuning $x/\kappa$: $n=0$ (black-dotted),
$n=0.7$ (green-dotted-dashed), $n=0.8$ (blue-dashed), and $n=0.9$ (red-solid)
with $\wp_{cL}=1mW$. In the inset, we plot the normalized output energy
$|\varepsilon_{outL+}/\varepsilon_{p}|^{2}$ vs. the ratio $n$.
\label{Fig5
\end{figure}
\begin{figure}[ptbh]
\centering \includegraphics[width=0.45\textwidth]{Fig6.eps}\caption{The
normalized right-hand output energy $|\varepsilon_{outR-}/\varepsilon_{p
|^{2}$ vs. the normalized frequency detuning $x/\kappa$: the parameters are
the same as in Fig. 4. In the inset, we plot the normalized output energy
$|\varepsilon_{outR-}/\varepsilon_{p}|^{2}+1$ vs. the ratio $n$.
\label{Fig6
\end{figure}
In Fig. 5-6, we plot the output power $|\varepsilon_{outL+}/\varepsilon
_{p}|^{2}$ and $|\varepsilon_{outR-}/\varepsilon_{p}|^{2}$ normalized to the
input probe field $\varepsilon_{p}$ respectively, versus the normalized
frequency $x/\kappa$ for different $n$. It can be seen clearly from Fig. 5-6
that the output energies $|\varepsilon_{outL+}/\varepsilon_{p}|^{2}$ and
$|\varepsilon_{outR-}/\varepsilon_{p}|^{2}$ get the maximum value at $x=0$ for
a certain value $n$. When $x=0$, the output normalized energies $|\varepsilon
_{outL+}/\varepsilon_{p}|^{2}$ and $|\varepsilon_{outR-}/\varepsilon_{p}|^{2}$
will increase with $n$ which is similar to the mechanical
oscillation $|\kappa\delta b_{+}/\varepsilon_{p}|^{2}$. This is because that
when $x=0$, the optomechanical effect will be strongest for a certain value
$n$ as discussed above. The curves of the output normalized energies
$|\varepsilon_{outL+}/\varepsilon_{p}|^{2}$ and $|\varepsilon_{outR-
/\varepsilon_{p}|^{2}$ almost have the same line shape, except that the output
normalized energy $|\varepsilon_{outL+}/\varepsilon_{p}|^{2}$ starts from 1
with the increase of $n$ for $x=0$ while the output normalized energy
$|\varepsilon_{outR-}/\varepsilon_{p}|^{2}$ starts from 0 (see the insets in
Fig. 5-6). This shows that the double-cavity optomechanical system will be reduced to
the standard one-cavity optomechanical model ($|\varepsilon_{outR-
/\varepsilon_{p}|^{2}=0$) when $n=0$. When $n$ increases up to 1, the
output normalized energies $|\varepsilon_{outL+}/\varepsilon_{p}|^{2}$ and
$|\varepsilon_{outR-}/\varepsilon_{p}|^{2}$ will increase approximately to
$1.6\times10^{5}$ (see the insets in Fig. 5-6). The reason for this is that,
the existing of blue-detuned driving field with $\omega_{d}-\omega_{0}=\omega_{m
$ will coherently enhance the oscillation of the MR (see Fig. 4),
leading to optomechanically induced amplification. Thus, we can realize the
optomechanically induced amplification for a resonantly injected probe in the double-cavity optomechanical system by appropriately
adjusting the ratio $n$ of the two strong field amplitudes
$\varepsilon_{c,d}$.
\section{Conclusions}
In summary, we have studied in theory a double-cavity optomechanical system
driven by a red sideband laser from one side and a blue sideband one from the other side. Our analytical and numerical results show that if adjusting the amplitude-ratio of the two driving fields $n>\sqrt{\gamma_{m}\kappa/2G^{2}}$ , the optomechanically induced
amplification for a resonantly incident probe (i.e., $\omega_{p}-\omega_{c}-\omega_{m}=0$) can be realized in this system. Typically, remarkable amplification can be obtained when $n\sim1$.
The reason for this is that, the Stokes processes in the blue-sideband driven cavity can generate
phonons in the mechanical elements, and these phonons will be further absorbed by the
Anti-Stokes processes in the red-sideband driven cavity. As a result, the optomechanical effect of the double-cavity
system is resonantly enhanced. In addition, the perfect optomechanically
induced transparency can be realized if we set the ratio $n=\sqrt
{\gamma_{m}\kappa/2G^{2}}$. Different from usual optomechanical induced transparency, this phenomenon is robust to mechanical dissipation, namely, the perfect transparency window can preserve even if the mechanical resonator has a relatively large decay rate $\gamma_{m}$. We expect that our study can be used to realize
signal quantum amplifier and light storage in the quantum information processes.
\begin{acknowledgments}
This work is supported by the National Natural Science Foundation of China (61378094). W. Z. Jia is supported by the National Natural Science Foundation of China under Grant No. 11347001 and the Fundamental Research Funds for the Central Universities 2682014RC21.
\end{acknowledgments}
\bigskip
|
\section{Introduction}
In the paper \cite{Berera14} we conducted direct numerical simulations (DNS)
of incompressible magnetohydrodynamic (MHD) turbulence without an imposed guide field.
The MHD equations for incompressible flow are
\begin{align}
\label{eq:momentum}
\partial_t \vec{u}&= - \frac{1}{\rho}\nabla P -(\vec{u}\cdot \nabla)\vec{u}
+ \frac{1}{\rho}(\nabla \times \vec{b}) \times \vec{b} + \nu \Delta \vec{u} \ , \\
\label{eq:induction}
\partial_t \vec{b}&= (\vec{b}\cdot \nabla)\vec{u}-(\vec{u}\cdot \nabla)\vec{b} + \eta \Delta \vec{b}\ , \\
\label{eq:incompr}
&\nabla \cdot \vec{u} = 0 \ \ \mbox{and} \ \ \nabla \cdot \vec{b} = 0 \ ,
\end{align}
where $\vec{u}$ denotes the velocity field, $\vec{b}$ the magnetic induction
expressed in Alfv\'{e}n units, $\nu$ the kinematic viscosity, $\eta$ the
resistivity, $P$ the pressure and $\rho=1$ the density.
We presented ensemble averaged data showing reverse spectral transfer (RST) of magnetic
helicity and measured decay exponents for initially maximally helical
magnetic fields.
In a particular set of simulations we decoupled the momentum equation
\eqref{eq:momentum} from the induction equation \eqref{eq:induction} and showed that the inverse
cascade of magnetic energy persists, while the reverse spectral transfer of magnetic
helicity disappears in the decoupled system, provided the initial velocity field was nonhelical.
Details of our simulations can be found in Table \ref{tbl:simulations}, where we specify values
for Reynolds numbers, resolution, diffusivities, runtime $t_{max}$, ensemble size
and wavenumber $k_0$ at which we set the peaks of the initial spectra.
\section{Verification of DNS code}
The 3D Orszag-Tang vortex was used for comparison against results by Mininni,
Pouquet and Montgomery \cite{Mininni06} and Morales, Leroy, Bos and Schneider \cite{Morales12}.
The initial conditions are
\begin{align}
{\vec{u}}&= (-2 \sin{y}, \ 2 \sin{x},\ 0) \nonumber \ , \\
{\vec{b}}&= \beta (-2 \sin{2y} + \sin{z}, \ 2 \sin{x} + \sin{z}, \ \sin{x} + \sin{y}) \ ,
\label{eq:OT_init}
\end{align}
where $\beta=0.8$ has been chosen according to \cite{Mininni06, Morales12}.
We calculate the total dissipation
\begin{equation}
\varepsilon_{tot}(t)= \nu \langle \vec{\omega}^2 \rangle + \eta \langle \vec{j}^2 \rangle \ ,
\end{equation}
and the maximum of the current density in real space $\mbox{max}|{\vec{j}}|$.
We used the same number of grid points and diffusivities as \cite{Morales12},
that is $\nu=\eta =0.01$ on $64^3$ points, $\nu=\eta =0.005$ on $128^3$ points
and $\nu=\eta =0.001$ on $256^3$ points.
Our results are in agreement with both above mentioned sources.
The maximum of the current density shows Reynolds number independent
exponential growth until $t=0.4$, in agreement with Morales \textit{et al.}~, while
Mininni \textit{et al.}~observe exponential growth up to $t=0.6$. After this initial period we
observe algebraic growth $\sim t^3$ in agreement with both sources. With
increasing Reynolds number the temporal maxima of $\mbox{max}|{\bf{j}}|$ are
achieved at later times, also in agreement with both
sources.
Our data shows very good agreement to a dataset obtained from
Morales \textit{et al.}~\cite{Morales12}, see Fig.~\ref{fig:Morales_comp}.
\begin{figure}[!t]
\begin{center}
\subfigure[ Evolution of the maximal current density]{
\includegraphics[width=0.48\textwidth]{maxcurrent_comp}
}
\subfigure[ Magnetic energy]{
\includegraphics[width=0.48\textwidth]{OT_Emag_comp}
}
\end{center}
\caption{Comparison to Morales \textit{et al.}~The symbols refer to our DNS, the lines to DNS by Morales \textit{et al.}}
\label{fig:Morales_comp}
\end{figure}
\section{Inertial range properties for the full MHD equations.}
Figure \ref{fig:inertial} shows magnetic helicity, magnetic energy and kinetic energy spectra
shortly after the onset of power-law decay (that is, at $t=1.3t_0$) for the largest Reynolds
number run H11. The magnetic helicity spectra have been multiplied by the wavenumber $k$
for dimensional reasons, and have been
shifted downwards in the figure to facilitate visual analysis. The inset shows
constancy of helicity flux $\Pi_H(k)$ for the wavenumber interval $21 \leqslant k \leqslant 33$,
indicating an inertial range for the magnetic helicity in the direct cascade region.
It can be seen in the figure that power-law scaling of $H_{mag}(k)$ extends over a larger interval,
showing the scaling $H_{mag}(k)\sim k^{-3.6}$, which is in agreement with
recent results on decaying 3D MHD turbulence \cite{Mueller12}.
The flux is $k$-dependent in the reverse spectral transfer
region, as shown in the paper in the inset of Fig.~1 for the lower $R_{\lambda}$-run H9 and
here in the inset of Fig.~\ref{fig:ic} for run H4a.
This is also in agreement with \cite{Mueller12}, who report the same behavior.
Our simulations support Kolmogorov scaling for the magnetic field as indicated
by the bar parallel to $E_{mag}(k)$.
\begin{figure}[!t]
\begin{center}
\includegraphics[width=0.7\textwidth]{inertial}
\end{center}
\caption{
Spectra for the largest Reynolds number run (H11). The solid line shows $E_{mag}(k)$,
the middle dashed line shows $E_{kin}(k)$ and the
bottom dotted line shows $kH_{mag}(k)$, which has been shifted for easier comparison.
The inset shows the flux of $H_{mag}(k)$, which is constant in the
higher $k$ region, indicating an inertial range. The straight lines indicate scaling regions
for $E_{mag}(k)\sim k^{-5/3}$ and $kH_{mag}(k)\sim k^{-2.6}$, which results in
$H_{mag}(k)\sim k^{-3.6}$.
}
\label{fig:inertial}
\end{figure}
\section{Evolution of the low wavenumber form of kinetic energy spectra for the full MHD equations.}
Figure \ref{fig:lowk} shows kinetic and magnetic energy spectra on
logarithmic scales at different times during the decay for run H9, which is the run with the
largest scale separation. Note that the slope of $E_{kin}(k)$ becomes flatter over time while
the slope of $E_{mag}(t)$ remains the same, that is $E_{mag}(k) \sim k^4$ at small $k$ prevails,
as indicated the diagonal line superimposed above $E_{mag}(k)$ in the
low wavenumber region. The slope of $E_{kin}(k)$
does \emph{not} change from $k^4$ to $k^2$, instead it continues to flatten over time.
\begin{figure}[!t]
\begin{center}
\includegraphics[width=0.7\textwidth]{E_magkin}
\end{center}
\caption{
(Color online) Evolution of kinetic (thin red (grey) lines) and magnetic (thick black lines) energy spectra over time for run H9.
Note how the slope of $E_{kin}(k)$ flattens with time, while the slope of $E_{mag}(k)$
remains unchanged.
}
\label{fig:lowk}
\end{figure}
\section{Further tests of linear RST}
We have conducted further tests in order to validate the linear RST found in our simulations. The velocity field should evolve as in hydrodynamic turbulence if
the Lorentz force term is omitted from the momentum equation. We have tested
that this is the case, see Fig.~\ref{fig:hydro}, this test clearly shows that
the velocity field evolves independently from the magnetic field.
\begin{figure}[!t]
\begin{center}
\subfigure[ Comparison to hydrodynamics]{
\label{fig:hydro}
\includegraphics[width=0.48\textwidth]{Ekin_test_dec}
}
\subfigure[ Results using the {\sc Pencil Code}]{
\label{fig:pencil}
\includegraphics[width=0.48\textwidth]{E_spec_pencil}
}
\end{center}
\caption{(a) The solid line shows $E_{kin}(t)$
from a hydrodynamic run, the symbols shows $E_{kin}(t)$ from a
simulation where the momentum equation was decoupled from the
induction equation, and the dashed line shows $E_{kin}(t)$ of
a full MHD run. The evolution of $E_{kin}(t)$ coincides for the
hydrodynamic run and the simulation of the decoupled case, showing that
in the decoupled case the velocity field evolves independently
of the magnetic field, as expected.
(b) Linear RST of magnetic energy using the {\sc Pencil Code} on $128^3$
grid points.
}
\end{figure}
\begin{figure}[!t]
\begin{center}
\subfigure[ Coupled case]{
\label{fig:ic}
\includegraphics[width=0.48\textwidth]{E_H_spectra_ins512}
}
\subfigure[ Decoupled case]{
\label{fig:ic_decoupled}
\includegraphics[width=0.48\textwidth]{E_H_spectra2}
}
\end{center}
\caption{(a) (Color online) Magnetic energy and helicity spectra showing reverse spectral transfer for a run on $512^3$ grid points (H4a).
The black (upper) lines refer to $E_{mag}(k)$, red (lower) lines refer to
$H_{mag}(k)$. Solid lines indicate one initial large eddy
turnover time, dotted and dash-dotted
lines refer to 2, 5 and 10 turnover times. The inset shows the
flux of magnetic helicity at one and two turnover times.
(b) Magnetic energy and helicity spectra for the decoupled case on $512^3$
grid points (Hd3). Upper (black) lines refer to $E_{mag}(k)$,
lower (red) lines to $H_{mag}(k)$. The inset shows the flux of magnetic helicity.
}
\end{figure}
The linear RST of magnetic energy was also found for slightly
compressible MHD using the
{\sc Pencil Code} \cite{Pencil}.
Figure \ref{fig:pencil} shows that after removing the Lorentz force term
in the momentum equation in the {\sc Pencil Code} we still observe RST of
magnetic energy.
Results for the linear RST of magnetic energy were found in simulations
using $256^3$ up to $1032^3$ collocation
points, but only the highest resolved results with the largest scale separation
are included in the paper \cite{Berera14}. Figure \ref{fig:ic_decoupled} shows
results from simulations
using $512^3$ collocation points (Hd3), where the peaks of the initial spectra had
been set at $k_0 = 5$. The linear RST is still visible, but concentrated at the
very lowest wavenumbers. In contrast, the simulation results
shown in the paper are obtained by setting the peaks of the initial spectra at
$k_0=23$, and RST can be seen up to $k=7$ in Fig. 3(a) of the paper.
Figure \ref{fig:ic_decoupled} also shows that RST is absent for the magnetic
helicity, as there is no increase of $H_{mag}(k)$ at the low wavenumbers and the flux
of magnetic helicity is negative. For comparison, the coupled case shows RST for both
$H_{mag}(k)$ and $E_{mag}(k)$ as can be seen in Fig.~\ref{fig:ic}, where the inset shows
positive flux of kinetic helicity in the low $k$ region.
Figure \ref{fig:dec_convergence} shows that our simulations have converged for
the time-step chosen. We do not see differences in the magnetic energy spectra
and the magnetic energy as we reduce the time-step, results are shown for
$256^3$ grid points.
\begin{figure}[!t]
\begin{center}
\subfigure[ Energy spectra]{
\includegraphics[width=0.4\textwidth]{emag_conv_spec}
}
\subfigure[ Time evolution of the magnetic energy]{
\includegraphics[width=0.4\textwidth]{emag_conv_stats}
}
\end{center}
\caption{
(a) Results for magnetic energy spectra from simulations run at different
time-steps on $256^3$ grid points. The spectra obtained from simulations using
different time-steps coincide.
(b) The evolution for the magnetic energy is the same for simulations using
different time-steps.
}
\label{fig:dec_convergence}
\end{figure}
\section{Importance of ensemble averaging}
\subsection{Decay exponents}
Figure ~\ref{fig:EA_exponents} illustrates the importance of ensemble
averaging for the measurement of the decay exponent $n_E$,
showing all realizations of H2 and the ensemble average.
It can clearly be seen how the decay exponents for single realizations
differ at later times, and how they deviate from the exponent measured at
early times. By contrast, the ensemble average follows the same straight line
for all times after the onset of power-law decay.
Similarly, a recent study of 2D hydrodynamic turbulence \cite{Mininni13}
using ensemble averaging showed a significant spread in the evolution of the
integral scale for different realizations and a somewhat less pronounded
spread in the total energy decay curves (see their Fig.~3).
The decay exponent $n_E$ measured
from individual realizations of H2 can range from 0.81-0.96 for $t>7t_0$,
where $t_0$ denotes the initial large eddy turnover time.
\begin{figure}[!t]
\begin{center}
\includegraphics[width=0.8\textwidth]{E_exp_single_red}
\end{center}
\caption{
Deviation of realizations (grey lines) from the ensemble average (black line)
of run H2.
Note that there is a clearly visible spread between realizations for
$t/t_0 \geq 7$.
}
\label{fig:EA_exponents}
\end{figure}
\subsection{Magnetic helicity for the decoupled case}
As mentioned in the paper \cite{Berera14}, some realizations of Hd5 appeared
to show RST, but the ensemble average did not confirm this. As can be seen
in Fig.~\ref{fig:EA_decIC}, the error on the ensemble averaged helicity
spectrum is large at $k=1$. Some realizations clearly show RST, some do not and
some realizations show negative helicity at $k=1$, which is why the value
does not appear on a logarithmic scale.
Taking data from one realization only could have lead to erroneous results,
claiming that either $H_{mag}(k)$ does show reverse spectral transfer, or that
$H_{mag}(k)$ becomes negative at $k=1$. The ensemble averaged result shows that
neither is the case. For $k=2$ there appears to be a hint of RST, but as can
be seen in Fig.~\ref{fig:EA_dec_err} the helicity spectrum at later time lies
within the error of the helicity spectrum at earlier time.
We would like to thank Jorge Morales for kindly providing us with a dataset of
simulation results using the Orszag-Tang vortex as initial condition.
\begin{figure}[H]
\begin{center}
\subfigure[ Helicity spectra of all realizations of the ensemble Hd5]{
\includegraphics[width=0.8\textwidth]{EA_Hspectra}
\label{fig:EA_decIC}
}
\subfigure[ Ensemble averaged helicity spectra for run Hd5]{
\includegraphics[width=0.8\textwidth]{EA_Hensemble}
\label{fig:EA_dec_err}
}
\end{center}
\caption{
(a) The black line shows the ensemble average of $H_{mag}(k)$, the grey lines
show $H_{mag}(k)$ for all realizations of run Hd5 at later time. Note the
large deviations between realizations at low $k$.
(b) The black line shows the ensemble average over the realizations in (a).
What appears as a hint of RST at $k=2$ lies within the error of the ensemble
average at earlier time (grey line).
}
\end{figure}
\begin{table}[H]
\centering
\begin{tabular}{ccccccccc}
Run id & $N^3$ & $k_{max}\eta_{mag}$ & $R_{\lambda}(0) $ & $\eta $ & $k_{0}$ & \# & $t_{max}/s$ \\%& $\tau(0)$ \\
\hline
\hline
H1 & $128^3$ & 1.30 & 28.69 & $0.009$ & 5 & 10 & 50 \\
H2 & $256^3$ & 2.42 & 32.27 & $0.008$ & 5 & 10 & 50\\
H3 & $256^3$ & 2.18 & 36.88 & $0.007$ & 5 & 10 & 50\\
H4 & $256^3$ & 2.08 & 43.03 & $0.006$ & 5 & 10 & 50 \\
H5 & $256^3$ & 1.80 & 51.64 & 0.005 & 5 & 10 & 50 \\
H6 & $256^3$ & 1.59 & 64.55 & 0.004 & 5 & 10 &50 \\
H7 & $256^3$ & 1.30 & 86.06 & $0.0031$ & 5 & 10 & 50 \\
H4a & $512^3$ & 1.38 & 43.03 & $0.002$ & 15 & 10 & 48 \\%&0.26\\
H8 & $512^3$ & 2.01 & 129.09 & $0.002$ & 5 & 10 & 7 \\%& 0.78\\
H9 & $1024^3$ & 1.38 & 74.84 & $0.00075$ & 23 & 10 & 6 \\%& 0.17\\
H10 & $528^3$ & 1.31 & 258.19 & $0.001$ & 5 & 10 & 50 \\%& 0.17\\
H11 & $1032^3$ & 1.38 & 645.47 & $0.0004$ & 5 & 5 & 12 \\%& 0.17\\
\hline
NH1 & $128^3$ & 1.29 & 28.69 & $0.009$ & 5 & 10 & 50 \\
NH2 & $256^3$ & 1.51 & 64.55 & $0.004$ & 5 & 10 & 10 \\
NH3 & $256^3$ & 1.26 & 86.06 & $0.003$ & 5 & 10 & 12 \\
NH4 & $512^3$ & 1.43 & 129.09 & $0.002$ & 5 & 10 & 10 \\
NH5 & $512^3$ & 1.30 & 172.13 & $0.0015$ & 5 & 10 & 7 \\
\hline
Hd1 & $256^3$ & 1.26 & 43.03 & $0.006$ & 5 & 10 & 5 \\%& 0.17\\
Hd2 & $256^3$ & 1.26 & 43.03 & $0.006$ & 5 & 10 & 5 \\%& 0.17\\
Hd3 & $512^3$ & 1.88 & 57.38 & $0.0045$ & 5 & 10 & 5 \\%& 0.17\\
Hd4 & $528^3$ & 1.95 & 57.38 & $0.0045$ & 5 & 10 & 5 \\%& 0.17\\
Hd5 & $1032^3$ & 1.39 & 28.06 & $0.001$ & 23 & 10 & 2 \\%& 0.17\\
\hline
NHd1 & $256^3$ & 1.26 & 43.03 & $0.006$ & 5 & 10 & 5 \\%& 0.17\\
NHd2 & $256^3$ & 1.26 & 43.03 & $0.006$ & 5 & 10 & 5 \\%& 0.17\\
NHd3 & $512^3$ & 1.88 & 57.38 & $0.0045$ & 5 & 10 & 5 \\%& 0.17\\
\end{tabular}
\caption{Specifications of simulations. H refers to an initially helical
magnetic field, NH to an initially non-helical magnetic field.
The additional letter d refers to the decoupled system, $R_{\lambda}(0)$
denotes the initial Taylor-scale Reynolds number, $\eta$ the
magnetic resistivity, $k_0$ the peak wavenumber of the initial energy spectra,
$k_{max}$ the largest resolved wavenumber, $\eta_{mag}$ the Kolmogrov microscale
associated with the magnetic field, \# the ensemble size and $t_{max}$ the
run time. The run NHd2 was
started from initial spectra following $k^2$ at low wavenumbers $k$, and
runs Hd2 and Hd4 were started using a maximally helical initial velocity field.
}
\label{tbl:simulations}
\end{table}
|
\section{Introduction}
In the past eleven years a series of charmoniumlike states $X, Y, Z$ has been announced at various experimental facilities (For recent reviews see, {\it e.g.}, Refs. \cite{bambrila,pr_navarra,navarra_mpla}). These exotic mesons are a class of hadrons that decay to final states that contain a heavy quark and a heavy antiquark but cannot be easily accommodated in the remaining unfilled states in the $c\bar{c}$ level scheme. One of the most interesting exotic states are the charged states, {\it e.g.} $Z(4430)$ and $Z_c(3900)$, that clearly have a more complex structure than $c\bar{c}$ pair, being natural candidates for molecular or tetraquarks states.
So far, most of the available experimental and theoretical investigations of the charmoniumlike exotic candidates have been focused on the spectrum and decays or the production in $e^+ e^-$ collisions. In order to decipher the nature of these states, more accurate data and new processes involving these states would be equally important. The search of $X,Y,Z$ charmoniumlike states by other production processes will not only confirm these observed states, but also will be useful to study their underlying structures. In the last years, the study of the production of exotic hadrons in proton - proton \cite{exoticpp}, heavy ion \cite{exotic_heavy} and photon - hadron \cite{z4430,y3940,zc3900,x3915} collisions was proposed by several authors. In particular, the results obtained in Refs. \cite{z4430,zc3900} for the photoproduction of charged charmoniumlike states $Z(4430)$ and $Z_c(3900)$ indicate a large enhancement of the cross section near the threshold, which could be considered a signature for the existence of these states in $\gamma p$ collisions. Unfortunately,
as the DESY - HERA $ep$ collider stopped its operations in 2007, the experimental analysis of these states in photon - hadron collisions was not possible and remains an open question.
In recent years we proposed the analysis of photon induced interactions in hadronic collisions as an alternative way to study the QCD dynamics at high energies
\cite{vicmag_upcs}.
The basic idea is that in this interactions the total cross section for a given process can be factorized in
terms of the equivalent flux of photons into the hadron projectile and the photon-photon or photon-target production cross section. The main advantages of using hadron - hadron collisions for
studying photon induced interactions are the high equivalent photon
energies and luminosities that can be achieved at existing accelerators (For a review see Ref. \cite{upcs}). In Ref. \cite{gama_prc} we estimated the production of some exotic states in $\gamma \gamma$ interactions considering $pp$ collisions at LHC energies and demonstrated that the experimental analysis of these states is, in principle, feasible. A similar conclusion was obtained in Ref. \cite{gama_bert} for $PbPb$ collisions.
In this paper we will demonstrate that the study of $\gamma p$ interactions
at the RHIC and LHC can also provide complementary and independent checks on the properties of the exotic states, and help to understand their underlying nature.
In this paper, we obtain, for the first time, an estimate of the photoproduction of the charmoniumlike states
$Y(3940)$, $X(3915)$, $Z(4430)$ and $Z_c(3900)$ in proton - proton collisions at the RHIC and LHC.
As we will show in the following, the prospect to observe these states is rather promising and thus an experimental analysis is strongly motivated.
This paper is organized as follows. In the next section we present a brief review of photon - hadron interactions in hadron - hadron collisions. In Section \ref{photo_exotic} we discuss the photoproduction of exotic charmoniumlike states and present the assumptions of the formalism used in our calculations. In Section \ref{resultados} we present our predictions for the
rapidity distribution and total cross sections for the
photoproduction of the exotic hadrons $Y(3940)$, $X(3915)$, $Z(4430)$ and $Z_c(3900)$ in $pp$ collisions at RHIC and LHC energies. Finally, in Section \ref{conc} we present a summary of our main conclusions.
\section{Photon - hadron interactions in $pp$ collisions}
Lets consider a hadron-hadron interaction at large impact parameter ($b > R_{h_1} + R_{h_2}$) and at ultra relativistic energies. In this regime we expect the dominance of the electromagnetic interaction.
In heavy ion colliders, the heavy nuclei give rise to strong electromagnetic fields due to the coherent action of all protons in the nucleus, which can interact with each other. In a similar way, it also occurs when considering ultra relativistic protons in $pp(\bar{p})$ colliders.
The photon stemming from the electromagnetic field
of one of the two colliding hadrons can interact with one photon of
the other hadron ($\gamma \gamma$ process) or can interact directly with the other hadron ($\gamma h$
process). For photon - hadron interactions the total
cross section for a given process can be factorized in terms of the equivalent flux of photons of the hadron projectile and the photon-target production cross section \cite{upcs}.
In the particular case of the photoproducion of a charmoniumlike state $H_c$ in a hadron-hadron collision, the total cross section is given by
\begin{eqnarray}
\sigma (h_1 h_2 \rightarrow p \otimes H_c h_3 ) = \sum_{i=1,2} \int dY \frac{d\sigma_i}{dY}\,,
\label{sighh}
\end{eqnarray}
where $h_1 = h_2 = p$, $\otimes$ represents a rapidity gap in the final state and $h_3 = p$ or $n$ depending if the interaction is mediated by a neutral ($\omega$) or a charged ($\pi^+$) particle, respectively. Moreover, ${d\sigma_i}/{dY}$ is the rapidity distribution for the photon-target interaction induced by the hadron $h_i$ ($i =1,2$), which can be expressed as
\begin{eqnarray}
\frac{d\sigma_i}{dY} = \omega \frac{dN_{\gamma/h_i}}{d\omega}\,\sigma_{\gamma h_j \rightarrow H_c h_3} (W_{\gamma h_j}^2) \,\,\,\,\,\,(i\neq j)\,.
\label{rapdis}
\end{eqnarray}
where $\omega$ is the photon energy, $\frac{dN_{\gamma/h}}{d\omega}$ is the equivalent photon flux, $W_{\gamma h}^2=2\,\omega \sqrt{s_{\mathrm{NN}}}$ and ${s_{\mathrm{NN}}}$ are the c.m.s energy squared of the
photon - hadron and hadron-hadron system, respectively.
In what follows we assume that the photon spectrum of a relativistic proton is given by \cite{Dress},
\begin{widetext}
\begin{eqnarray}
\frac{dN_{\gamma/p}(\omega)}{d\omega} = \frac{\alpha_{\mathrm{em}}}{2 \pi\, \omega} \left[ 1 + \left(1 -
\frac{2\,\omega}{\sqrt{s_{NN}}}\right)^2 \right] .
\left( \ln{\Omega} - \frac{11}{6} + \frac{3}{\Omega} - \frac{3}{2 \,\Omega^2} + \frac{1}{3 \,\Omega^3} \right) \,,
\label{eq:photon_spectrum}
\end{eqnarray}
\end{widetext}
where $\Omega = 1 + [\,(0.71 \,\mathrm{GeV}^2)/Q_{\mathrm{min}}^2\,]$, $Q_{\mathrm{min}}^2= \omega^2/[\,\gamma_L^2 \,(1-2\,\omega /\sqrt{s_{NN}})\,] \approx (\omega/
\gamma_L)^2$ and $\gamma_L$ is the Lorentz factor.
The experimental separation for such events is relatively easy, as photon emission is coherent over the hadron and the photon is colorless we expect the events to be characterized by intact recoiled hadron (tagged hadron) and the presence of a rapidity gap (For a detailed discussion see \cite{upcs}). Moreover, the detection of the neutron in the final state can be useful to separate the events associated to the production of a charged charmoniumlike state. Finally, it is important to emphasize that as the photon spectrum decreases with energy approximately like $1/\omega$, the main contribution for the total cross section, Eq. (\ref{sighh}), comes from the small $\omega$ behaviour of the photon spectrum where we are probing the photon - hadron cross section at low values of the center-of-mass energy.
\begin{figure}[t]
\begin{tabular}{cc}
\includegraphics[scale=0.28]{zc3900.eps} &
\includegraphics[scale=0.25]{zc3900_pom.eps}
\end{tabular}
\caption{(a) The $pp \rightarrow p Z_c^+ n \rightarrow p J/\Psi \pi^+ n$ process through the $\pi^+$ exchange. (b) The $pp \rightarrow p J/\Psi \pi^+ n$ process through the Pomeron exchange. }
\label{diagrama}
\end{figure}
\section{Photoproduction of exotic charmoniumlike states}
\label{photo_exotic}
The main input in our calculations is the photon - hadron cross section for the production of the of exotic charmoniumlike state $\sigma_{\gamma p \rightarrow H_c}$.
Currently the unique available predictions in the literature \cite{z4430,y3940,zc3900,x3915} were obtained considering an effective Lagrangian approach combined with the vector meson dominance (VMD) assumption \cite{vdm_3}. In this approach the interaction is described in terms of a meson exchange, which is neutral/charged for the production of neutral/charged exotic charmoniumlike states. In what follows we briefly review the main assumptions of the approach proposed in Refs. \cite{z4430,y3940,zc3900,x3915}, with particular emphasis in the production of the $Z_c(3900)^+$. The results for the other exotic states are obtained in a similar way.
Considering the approach proposed in Ref. \cite{zc3900} the photoproduction of the $Z_c(3900)^+$ in $pp$ collisions is described by the diagram presented in Fig. \ref{diagrama}(a), where we also represent the decay of the exotic meson in a $J/\Psi + \pi^+$ final state, which we assume to be the dominant channel. The basic idea is that the photon stemming from the electromagnetic field of one of the incident protons fluctuates into a $J/\Psi$ which interacts with the other proton through the $\pi^+$ exchange producing a neutron $n$ and a $Z_c(3900)^+$ state which decays in the $J/\Psi + \pi^+$ system.
Assuming the VMD model to describe the coupling between the intermediate vector meson and the photon, an effective Lagrangian to describe the coupling between the pion and the nucleons and another describing the $Z_c J/\Psi \pi$ coupling, the squared amplitude for the process $\gamma p \rightarrow Z_c^+ n$ can be expressed by (See Ref. \cite{zc3900} for details)
\begin{eqnarray}
|{\cal{M}}|^2 & = & {\cal{C}} \frac{-q^2(q^2 - M_{Z_c}^2)^2}{(q^2 - m_{\pi}^2)^2} F_{\pi NN}^2 F_{Z_cJ/\Psi \pi}^2
\end{eqnarray}
where $q$ is the four momentum of the pion exchanged and
\begin{eqnarray}
{\cal{C}} = \left(\sqrt{2} g_{\pi NN} \frac{g_{Z_cJ/\Psi \pi}}{M_{Z_c}} \frac{e}{f_{J/\Psi}}\right)^2
\end{eqnarray}
with the couplings $g_{\pi NN}$, ${e}/{f_{J/\Psi}}$ and $g_{Z_cJ/\Psi \pi}$ being respectively determined assuming that $g_{\pi NN}^2/4 \pi = 14$, by the decay $J/\Psi \rightarrow e^+ e^-$ and by the decay width of $Z_c \rightarrow J/\Psi \pi$. In what follows we assume $\Gamma[Z^+ \rightarrow J/\Psi \pi^+] = 29$ MeV as calculated in \cite{maiani}. Moreover, the form factors $F_{\pi NN}$ and $F_{Z_cJ/\Psi \pi}$ are given by
\begin{eqnarray}
F_{\pi NN} = \left(\frac{\Lambda_{\pi}^2 - m_{\pi}^2}{\Lambda_{\pi}^2 - q^2}\right) \,, \,\,\,
F_{Z_cJ/\Psi \pi} = \left(\frac{\Lambda_{Z_c}^2 - m_{\pi}^2}{\Lambda_{Z_c}^2 - q^2}\right)
\end{eqnarray}
where we assume $\Lambda_{\pi} = 0.7$ GeV and $\Lambda_{Z_c} = M_{J/\Psi}$ for the cut-offs. The resulting $\gamma p$ cross section is strongly enhanced close to the threshold \cite{zc3900}.
\begin{figure}[t]
\includegraphics[scale=0.35]{gamma_p.eps}
\caption{(Color online) Energy dependence of the photoproduction cross sections. The signal (Meson + $\tt I\! P$) and background ($\tt I\! P$) contributions are presented separately. }
\label{gamap}
\end{figure}
This formalism can be directly extended to take into account the $Z_c$ decay in the $J/\Psi + \pi$ final state. However, as this final state can also be produced in a $\gamma p$ interaction through the Pomeron exchange, we should include this background in our calculations. In Fig. \ref{diagrama}(b) we present a typical diagram for the background contribution. Following Ref. \cite{ln} we assume that the Pomeron behaves like an isoscalar photon with $C=+1$. The corresponding amplitude has a Regge-like energy dependence $\propto s^{\alpha_{\tt I\! P}(t) - 1}$, where $\alpha_{\tt I\! P}(t) = 1 + \Delta + \alpha^{\prime}t$ is the Pomeron trajectory, $\Delta$ is the Pomeron intercept, $\alpha^{\prime} = 0.25$ GeV$^{-2}$ and $t$ is the exchanged Pomeron momentum squared. In Fig. \ref{gamap}(a) we present our predictions for the energy dependence of the $\gamma p \rightarrow J/\Psi \pi n$ cross section considering the signal ($\pi + \tt I\! P$) and background ($\tt I\! P$) contributions. We obtain that the resulting cross section for the $\gamma p \rightarrow J/\Psi \pi n$ process is dominated by the signal for $W_{\gamma p} \le 20$ GeV, in agreement with the results presented in Ref. \cite{zc3900}.
\begin{figure}[t]
\includegraphics[scale=0.3]{Zc3900_rhic200.eps} \\
\includegraphics[scale=0.3]{Zc3900_lhc14.eps}
\caption{(Color online) Rapidity distribution for the photoproduction of a $J/\Psi + \pi$ final state in $pp$ collisions at RHIC ($\sqrt{s} = 0.2$ TeV) and LHC ($\sqrt{s} = 14$ TeV) energies. The solid lines represent the background associated to the Pomeron exchange for two values of the intercept $\Delta$. The dashed lines represent the sum of the background with the signal associated to the $\gamma p \rightarrow Z_c(3900)^+ n \rightarrow J/\Psi \pi n$ interaction through $\pi$ exchange.}
\label{zc3900_14}
\end{figure}
\begin{figure}[ht]
\includegraphics[scale=0.3]{Y3940_rhic200.eps} \\
\includegraphics[scale=0.3]{Y3940_lhc14.eps}
\caption{(Color online) Rapidity distribution for the photoproduction of a $J/\Psi + \omega$ final state in $pp$ collisions at RHIC ($\sqrt{s} = 0.2$ TeV) and LHC ($\sqrt{s} = 14$ TeV) energies. The solid lines represent the background associated to the Pomeron exchange ($\Delta$ = 0.08). The dashed lines represent the sum of the background with the signal associated to the $\gamma p \rightarrow Y(3940)+ p \rightarrow J/\Psi \omega p$ interaction through $\omega$ exchange.}
\label{y3940}
\end{figure}
A similar analysis can also be performed for the photoproduction of the $Z(4430)^+$. The main differences are that we assume that the coupling of the photon occurs with a $\Psi^{\prime}$ instead of the $J/\Psi$ and that the dominant decay channel of the $Z(4430)^+$ is in a $\Psi^{\prime} + \pi^+$ system. Following Ref. \cite{z4430}
we also assume a $Z(4430)$ with $J^P = 1^-$ quantum numbers. However, we expect to obtain similar results for a state with $J^P = 1^+$. As in the $Z_c(3900)^+$ case, we obtain in Fig. \ref{gamap}(b) that the $\sigma (\gamma p \rightarrow \Psi^{\prime} \pi n)$ is dominated by the signal at small values of the $\gamma p$ center-of-mass energy, with a strong enhancement close to the threshold, reproducing the results presented in Ref. \cite{z4430} for $\Lambda_{Z \rightarrow \Psi^{\prime} \pi} = 45$ MeV and $g_{Z \Psi^{\prime} \pi} = 2.365$ GeV$^{-1}$. Finally, for the photoproduction of the neutral charmoniumlike states
$Y(3940)$ and $X(3915)$ we assume that these states are $0^{++}$ and $2^{++}$ states, that
the cross sections for the signals are described in terms of a $\omega$ exchange and that these states decay into a $J/\Psi + \omega$. It is important to emphasize that distinctly from the production of charged states, for the case of neutral states, the proton target remains intact in the final state. Consequently, the final state for the photoproduction of these states in $pp$ collisions is characterized by two intact protons. Our predictions for $\sigma (\gamma p \rightarrow J/\Psi \pi p)$ are presented in Figs. \ref{gamap}(c) and (d) for $\Gamma[Y(3940) \rightarrow J/\Psi \omega] = 34$ MeV and $\Gamma[X(3915) \rightarrow J/\Psi \omega] = 5.10$ MeV and are similar to those obtained in Refs. \cite{y3940,x3915}. It is important to emphasize that for these states the total cross section do not have a resonancelike structure close to the threshold, with the signal being larger than the background for $W_{\gamma p} \le 50 \,(20)$ GeV for the $Y(3940)$ ($X(3915)$) photoproduction.
\section{Results}\label{resultados}
Lets initially calculate the rapidity distribution and total cross section for the photoproduction of the charged charmoniumlike states $Z_c(3900)$ and $Z(4430)$.
The distribution on rapidity $Y$ of the hadron $H_c$ of mass $M_{H_c}$ in the final state can be directly computed from Eq. (\ref{sighh}), by using its relation with the photon energy $\omega$, i.e. $Y\propto \ln \, ( \omega/M_{H_c})$. Explicitly, the rapidity distribution is written down as,
\begin{widetext}
\begin{eqnarray}
\frac{d\sigma \,\left[h_1 h_2 \rightarrow p \otimes H_c + n \right]}{dY} = \left[\omega \frac{dN}{d\omega}|_{h_1}\,\sigma_{\gamma h_2 \rightarrow H_c + n}\left(\omega \right)\right]_{\omega_L} + \left[\omega \frac{dN}{d\omega}|_{h_2}\,\sigma_{\gamma h_1 \rightarrow H_c + n}\left(\omega \right)\right]_{\omega_R}\,
\label{dsigdy2}
\end{eqnarray}
\end{widetext}
where $\omega_L \, (\propto e^{-Y})$ and $\omega_R \, (\propto e^{Y})$ denote photons from the $h_1$ and $h_2$ hadrons, respectively. As the photon fluxes, Eq. (\ref{eq:photon_spectrum}), have support at small values of $\omega$, decreasing exponentially at large $\omega$, the first term on the right-hand side of the Eq. (\ref{dsigdy2}) peaks at positive rapidities while the second term peaks at negative rapidities. Consequently, given the photon flux, the study of the rapidity distribution can be used to constrain the photoproduction cross section for a given energy. Moreover, the total rapidity distributions for $pp$ collisions will be symmetric about midrapidity ($Y=0$).
In Fig. \ref{zc3900_14} we present our predictions for the rapidity distribution for the production of a $J/\Psi + \pi$ final state in $pp$ collisions at RHIC ($\sqrt{s} = 0.2$ TeV) and LHC ($\sqrt{s} = 14$ TeV) energies considering the diagrams presented in Figs. \ref{diagrama} (a) and (b). The cross section for the background, associated to the Pomeron exchange, has been estimated considering two different values for the Pomeron intercept $\Delta = 0$ and 0.08, which determines its energy dependence. At RHIC energy these two predictions for the Pomeron contribution are almost identical, while at LHC they differ by almost 10 $\%$ at $Y = 0$. We predict that the signal dominates the rapidity distributions for large values of the rapidity $|Y|$, where we are probing the $\gamma p$ cross section at small values of the c.m. energy. The resonancelike structure present in the $\gamma p$ cross section, which is a signature of the $Z_c^+$ production, can be directly probed at $|Y| \geq 2.0 \, (5.5)$ at RHIC (LHC) energy. A similar behaviour is predicted in the rapidity distribution for the photoproduction of the $Z(4430)^+$.
Our predictions for the photoproduction of the neutral charmoniumlike state $Y(3940)$ decaying into a $J/\Psi + \omega$ state are presented in Fig.
\ref{y3940}. In this case we predict that the signal is larger than the background at all values of rapidity. In particular, we predict that the signal is larger by a factor $\ge 5$ at $Y = 0$ . A similar conclusion is reached for the $X(3915)$.
In Table \ref{tab1} we present our predictions for the total cross sections for the different final states discussed above. In particular, we present the predictions for the background contributions as well as for the signals which take into account the production of different exotic charmoniumlike states. We predict that the cross sections are of $\cal{O}$(nb) for RHIC and LHC energies. It is important to emphasize that similar numbers are obtained for the Tevatron energy. For the charged states the signal is larger than the background by a factor $\ge 1.6$, being about three for the RHIC energy. In order to estimate the number of events in $pp$ collisions at LHC, we assume the design luminosity ${\cal L} = 10^7$ mb$^{-1}$ s$^{-1}$. Consequently, we predict ${\cal{O}}(10)$ events/second for these processes.
\section{Conclusions} \label{conc}
During the last decade a rich spectroscopy of charmoniumlike states that do not fill into the remaining unassigned levels for $c\bar{c}$ charmonium states has emerged. However, several questions still remain to be answered. In this paper we have proposed, for the first time, the study of the photoproduction of exotic charmoniumlike states in $pp$ collisions at RHIC and LHC energies. Considering an effective Lagrangian description for the photon - hadron interaction we have derived an estimate of the production rates for these particles. Our results show that the resulting cross sections for the $Y(3940)$, $X(3915)$, $Z(4430)$ and $Z_c(3900)$ photoproduction at the RHIC energy are at 1 nb level, while at LHC are larger by roughly one order of magnitude.
Taking into account the design luminosity at the LHC the number of events has been estimated. Our results indicate that these four exotic states could be copiously produced. Consequently, we believe that study of these states at RHIC and LHC is feasible and will provide valuable information on hadron spectroscopy as well as hadron interactions.
Finally, measurements of these exotic states at RHIC and LHC will supplement the results obtained at $e^+e^-$ colliders, and thus allow to explore the nature of the states.
\begin{widetext}
\begin{table}[t]
\begin{center}
\begin{tabular} {|c|c|c|c|c|c|}
\hline
\hline
Reaction & Ressonance & Contribution & $\sigma$ [nb] ($\sqrt{s} = 0.2$ TeV)
& $\sigma$ [nb] ($\sqrt{s} = 7$ TeV) & $\sigma$ [nb] ($\sqrt{s} = 14$ TeV) \\
\hline
\hline
$\sigma(pp \rightarrow pJ/\Psi\pi n)$ & -- & $\tt I\! P$ & 1.15 & 8.18 -- 9.64 & 10.33 -- 12.65 \\
&$Z_c(3900)$ & $\tt I\! P + \pi$ & 3.83 & 14.13 -- 15.52 & 16.89 -- 19.12 \\
\hline
$\sigma(pp \rightarrow p\Psi^{\prime}\pi n)$ & -- & $\tt I\! P$ & 2.60 & 18.15 -- 21.32 & 22.87 -- 27.93 \\
&$Z(4430)$ & $\tt I\! P + \pi$ & 7.33 & 29.26 -- 32.41 & 35.21 -- 40.23 \\
\hline
$\sigma(pp \rightarrow pJ/\Psi\omega p)$ & -- & $\tt I\! P$ & 0.84 -- 0.90 & 5.90 -- 7.75 & 7.42 -- 10.17 \\
& $X(3915)$ & $\tt I\! P + \omega$ & 1.88 -- 1.98 & 11.31 -- 14.53 & 14.08 -- 18.88 \\
\hline
$\sigma(pp \rightarrow pJ/\Psi\omega p)$ & -- & $\tt I\! P$ & 1.33 & 12.73 -- 15.35 & 16.35 -- 20.54 \\
& $Y(3940)$ & $\tt I\! P + \omega$ & 12.62 & 74.28 -- 85.93 & 92.58 -- 111.19 \\
\hline
\hline
\end{tabular}
\end{center}
\caption{Total cross sections for the photoproduction of different final states in $pp$ collisions at RHIC and LHC energies considering the sum of the signal associated to the photoproduction of an exotic charmoniumlike state, produced by a $\pi$ or $\omega$ exchange, and the background contribution associated to the Pomeron ($\tt I\! P$) exchange. For comparison the magnitude of the background contribution is presented separately.}
\label{tab1}
\end{table}
\end{widetext}
\section*{Acknowledgments}
The authors are grateful to A. Martinez Torres, K. Kemchandani, F. S. Navarra and M. Nielsen for discussions.
This work has been supported by CNPq and FAPERGS, Brazil.
|
\section{nicht vergessen}
\section{Introduction}
The Standard Model of elementary particles (SM) is very successful on the phenomenological level. The outcome of (almost) any particle physics experiment can be predicted accurately within this model, and, where not, by some straightforward extension. For example, one may introduce right handed neutrinos to account for tiny neutrino masses\cite{nuemix}.
Nevertheless, it is widely believed that the SM is only an effective low-energy theory valid below a certain energy scale $\Lambda_r$, which is supposed to be of the order of 1-10 TeV. This view is based on the fact that the SM has many unknown parameters and one rather mysterious component, the so-called Higgs field, which is needed for the spontaneous symmetry breaking (SSB) taking place in the model. Within the Higgs sector the most challenging part is the set of Yukawa couplings to fermions, which comprises the majority of the unknown parameters of the SM Lagrangian.
In recent papers a microscopic model of the Higgs mechanism has been developed\cite{laa,lamm,lamm1}, and also an idea how the quark and lepton states arise in that model. This concept will be used in the present article in an attempt to determine the fermion masses and mixings.
At first sight, the spectrum of quarks and leptons seems difficult to explain, because it extends over many orders of magnitude, starting from the neutrinos with their tiny masses below 1 eV, passing over to the 'everyday life' particles e, u and d with masses of order $10^6$\ eV, proceeding to muon and strange-quark (about $10^8$eV), ascending to charm, $\tau$ and bottom, which have masses of order $10^9$eV, and finishing with the top quark, whose mass of 1.7$\times 10^{11}$eV lies suspiciously close to the SSB scale $\Lambda_F$, the value of the Higgs mass and to twice the W-mass. For the neutrinos it is reasonable to believe that their mass might be some kind of higher order effect\cite{radi}, is protected by symmetry or generated by a variant of the popular seesaw mechanism\cite{seesaw,seevalle}. Other approaches to explain the hierarchy in the particle masses consider textures\cite{ramond} like 'democratic' mass matrices\cite{democ} with identical entries. These show the desirable feature that after diagonalization there is one very heavy particle (the top quark), and the rest have small masses.
Unfortunately, a physical understanding of the underlying dynamics responsible for these effects is still lacking. For example, in (supersymmetric) grand unified theories fermion masses essentially remain free parameters. Furthermore, those models usually introduce many more additional degrees of freedom without much ambition to determine them from first principles. The point is that theories of that kind only extrapolate and extend the symmetries observed at low energies to small distances and that there is a strong amount of arbitrariness in this procedure. In my opinion it is obvious that a physical understanding of the masses and mixings is only possible in a microscopic theory. Superstring theories seem to offer such an understanding. However, although 'in principle' able to determine the masses as energies of string excitations, to my knowledge they have not come up with definite and verifiable predictions.
The present study is devoted to partially fill this gap. Some of the above mentioned 'textures' will reappear in the sections below. For example, neutrinos are indeed protected by the symmetries of new interactions. Furthermore, a kind of democratic texture will be derived which makes the top quark the heaviest fermion. More precisely, it arises from a symmetry breaking contribution modified by a Dzyaloshinskii-Moriya component\cite{dmgut} in such a way that all entries of the mass matrix effectively get identical contributions. The appearance of this modification turns out to be a reflection of the $SU(2)_L$ gauge symmetry (breaking) on the microscopic level.
A related question is how the mixing between the generations can be understood. Most of the mixing angles are now known with a reasonable accuracy\cite{ckm,nuemix}. In particular, there is a hierarchy in the mixing matrix for the quark sector, but not in the lepton sector. Approaching the mixing problem in the present model I will be able to give some preliminary results mainly for the neutrinos and also set up the environment to derive the CKM mixing angles. In the course of the discussion an explanation will be given, of why the quark mixings are naturally 'small' and the lepton mixings naturally 'large'.
\section{Quarks and Leptons as Isospin Excitations of a Tetrahedral Shubnikov Group}
The Higgs doublet of the Standard Model can be parametrized as
\begin{eqnarray}
\Phi=\frac{1}{\sqrt{2}}
\begin{pmatrix}
i(\pi_x-i \pi_y) \\
\sigma -i\pi_z
\end{pmatrix}
\label{hig77}
\end{eqnarray}
where $\sigma = \Lambda_F +\phi$ acquires a vacuum expectation value $\langle \sigma \rangle =\Lambda_F=\sqrt{\frac{\mu^2}{\lambda}}=246$GeV through the form of the potential
\begin{eqnarray}
V(\Phi)&=&-\mu^2 \Phi^+ \Phi + \lambda (\Phi^+\Phi)^2= -\frac{1}{2}\mu^2 (\sigma^2 +\vec\pi^2)+\frac{1}{4}\lambda (\sigma^2 +\vec\pi^2)^2 \nonumber \\
&=&\frac{1}{4}\lambda [-\Lambda_F^4+4\Lambda_F\phi\vec\pi^2+4\Lambda_F^2\phi^2+4\Lambda_F\phi^3+\phi^4+\vec\pi^4+2\phi^2\vec\pi^2]
\label{hipo}
\end{eqnarray}
and $\vec \pi =(\pi_x,\pi_y,\pi_z)$ gets 'eaten' by the longitudinal modes of the afterwards massive W-bosons. This can be made explicit by a SU(2) gauge transformation of the form\cite{technireview}
\begin{eqnarray}
U=\exp (\frac{i\vec \tau \vec \pi}{2\Lambda_F} )
\label{hipoga}
\end{eqnarray}
which formally removes $\vec \pi$ from the Higgs doublet.
\begin{figure}
\begin{center}
\epsfig{file=gut-tetraeder4.eps,height=6.0cm}
\caption{The local ground state of the model, living in a 3-dimensional internal $R^3$ space. Shown are the corner points (small circles) of the internal tetrahedron, which can be represented by their coordinate vectors $\vec r_i$. The origin of coordinates is taken to be the center of the tetrahedron, and is identical to the base point of the fiber in Minkowski space.
On each corner point $i=1,2,3,4$ there is an axial spin vector $\vec \pi_i$, pointing in the same radial direction as $\vec r_i$.}
\nonumber
\end{center}
\end{figure}
The present calculation is motivated by a Nambu-Jona-Lasinio (NJL) type of interpretation of the SSB mechanism. This heuristic ansatz will then be vigorously extended to a new, microscopic picture of the SM particle system. The starting point is an isospin pair $\psi=(U,D)$ of Dirac fermions distantly similar as in technicolor models\cite{techni1,techni2,techni3,techni4}, however without a technicolor quantum number. Rather we shall assume that the pairing mechanism is due to exchange interactions and strong correlations between fermions, effects which in many body physics are known to be responsible for SSB in superconductors and (anti)-ferromagnets. In contrast to solid state physics we do not consider these effects in physical space, but attribute them to arise from an independent dynamics which is active in the internal spaces. It is this dynamics which will allow us to put hand on the values of the fermion masses.
The main idea is that (weak) isospin arises from a nonrelativistic internal $R^3$ space much like ordinary spin arises from physical space. In other words, the internal space is assumed to possess a rotational SO(3)-symmetry for which the doublet $\psi=(U,D)$ serves as an (internal) Pauli spinor with an initial internal SU(2) spin symmetry. These internal spins are assumed to undergo interactions in the internal space which can be described by internal Heisenberg spin interactions which are formally similar to those describing spin interactions in solids.
The geometrical picture is that the world is a fiber bundle over Minkowski space with fibers given by internal $R^3$ spaces, and that within these fibers physical processes take place, which can be described by a higher dimensional quantum electrodynamics.
This idea was worked out in ref. \cite{laa} and the interested reader is referred to that paper for more details. It is further assumed that at high temperatures there is a symmetric phase in which the internal spins are distributed randomly in the fibers, giving rise to a local internal SU(2) symmetry of the Lagrangian, local in the sense that on each site in each fiber the spins may be rotated independently.
When the temperature of the universe decreases from big bang energies to TeV values the internal magnetic interactions within the fiber lead to the frustrated\cite{frust} tetrahedral structure shown in fig. 1, and when it falls below the Fermi scale all the tetrahedrons over Minkowski space align as in fig. 2, a process which in ref.\cite{laa} was claimed to be the microscopic origin of the Higgs mechanism. A pairing process for the formation of the Higgs particle has also been described in that paper.
The configuration fig. 1 is the starting point of the present calculation, because it is considered as the {\it local} ground state of the system. In other words, it is assumed that in each of the 3-dimensional internal $R^3$ fibers there is a discrete tetrahedral structure and that the internal dynamics is such that spin vectors arrange themselves according to this internal tetrahedral configuration. The tetrahedron itself has the tetrahedral group $S_4$ as point group symmetry. However, due to the pseudovector property of the internal spin vectors the whole system loses its reflection symmetries and obtains instead the Shubnikov symmetry group $A_4 + S ( S_4 - A_4)$\cite{shub,white,borov}, where S is the internal time reversal operation and $A_4$ is the subgroup of $S_4$ which does not contain reflections. Note that S itself does not belong to the Shubnikov group, and also the internal reflections do not. The Shubnikov group is chiral, the configuration with opposite chirality being given when the 4 spin vectors would point inwards instead of outwards. Before the formation of the chiral tetrahedron the internal spins U and D, which according to eq. (\ref{njl09y}) are the building blocks of the spin vectors $\vec \pi_i$, can freely rotate and thus there is an internal spin SU(2) symmetry group, which however is broken to $A_4 + S ( S_4 - A_4)$ when the chiral tetrahedron is formed.
With respect to (external) Lorentz symmetry both U and D can appear as lefthanded or righthanded objects, so that one may in fact consider separately a $SU(2)_L$ for the lefthanded and $SU(2)_R$ for the righthanded objects. Before the advent of the chiral tetrahedrons and of the gauge bosons the Higgs sector of the SM is symmetric under $SU(2)_L \times SU(2)_R$, and a Nambu-Jona-Lasinio (NJL) model with this symmetry may we formulated.
\begin{figure}
\begin{center}
\epsfig{file=gut-tetrad6.eps,height=4.4cm}
\bigskip
\caption{The global ground state of the model after SSB consists of an aligned system of chiral tetrahedrons over Minkowski space (the latter represented by the long arrow). R is the magnitude of one tetrahedron and r the distance between two of them. Note that contrary to what is drawn here, the tetrahedra extend into internal space alone, not into Minkowski space. Before the SSB the chiral tetrahedrons are oriented randomly (not shown) and there is a corresponding {\it local} $SO(3)$ symmetry, because each rigid tetrahedron can be rotated freely and independently from the others.}
\nonumber
\end{center}
\end{figure}
The NJL philosophy operates as follows: the SSB is induced by formation of bound states and condensates of the fermion doublet $\psi=(U,D)$. Namely, the quadratic part
\begin{eqnarray}
V_2(\Phi)=-\mu^2 \Phi^+ \Phi= -\frac{1}{2}\mu^2 (\sigma^2 +\vec\pi^2)
\label{hipo700}
\end{eqnarray}
of the potential eq. (\ref{hipo}) is equivalent to a NJL interaction of the form
\begin{eqnarray}
V_{NJL}=G_{NJL}[ (\bar \psi \psi)^2+(\bar \psi i \gamma_5 \vec \tau \psi)^2]
\label{njl09}
\end{eqnarray}
where $G_{NJL}$ denotes the NJL coupling strength, which in the SSB regime, where $V_2(\Phi)<0$, must be negative as well. (Note that $\bar \psi i \gamma_5 \vec \tau \psi$ is always real.) In fact using the method of auxiliary fields one can show $V_2(\Phi)=V_{NJL}$ provided one chooses
\begin{eqnarray}
\sigma &=&-2G_{NJL}\bar \psi \psi \nonumber \\
\vec\pi &=&-2G_{NJL}\bar \psi i \gamma_5 \vec \tau \psi
\label{njl09y}
\end{eqnarray}
and thus obtain a sigma model from the original NJL potential (\ref{njl09}).
For later use I will rewrite eq. (\ref{njl09}) as
\begin{eqnarray}
V_{NJL}=-J_F [ (\frac{\bar \psi \psi}{\mu^3})^2+(\frac{\bar \psi i \gamma_5 \vec \tau \psi}{\mu^3})^2]
\label{njl229}
\end{eqnarray}
so that in brackets there are dimensionless quantities and $J_F=-G_{NJL}\mu^6$ has the dimension 4 of an energy density.
This equation has been interpreted in ref. \cite{laa} to describe the dynamics of interacting chiral internal spin vectors $\vec \pi$, normalized to the SSB scale $\mu$.
Here 'chiral' refers both to internal and physical space, because $\gamma_5$ is a building block for chiral objects in physical space, while the appearance of (internal) Pauli matrices $\vec \tau$ signals chiral objects in (internal) space. Moreover, in the framework of the internal Heisenberg spin theory $J_F$ is to be interpreted as the internal exchange energy density corresponding to certain exchange integrals over internal $R^3$ space to be described later. In the SSB regime one has $J_F>0$ (because of $G_{NJL}<0$) corresponding to a ferromagnetic interaction. This interaction accounts for neighbouring tetrahedrons aligning themselves over Minkowski space as shown in fig. 2.
There is a slight complication on these considerations, because at SSB energies after redefinition of $\sigma = \Lambda_F +\phi$, the $\vec\pi$-$\vec\pi$ interaction eq. (\ref{hipo700}) seems to disappear, because the sum of terms $\sim \vec\pi^2$ vanishes in the potential $V(H)$ as made explicit by the last of eqs. (\ref{hipo}). However, when the $\vec \pi$ triplet is absorbed as the longitudinal mode of the $\vec W$-boson the internal Heisenberg spin interaction reappears as part of the mass term $m_W^2 W_\mu W^\mu$. The alignment of tetrahedrons in fig. 2 will then experience a modification which is dictated by the gauge symmetry of the fiber bundle formed by all tetrahedrons. As a consequence the ferromagnetic Heisenberg interaction has to be modified by a Dzyaloshinskii-Moriya component\cite{dmgut,dmz1,dmz2} in this regime. Details will be given in section 5.
Next I want to extent the view to small distances and high energies. At high energies, there is no SSB and instead of the negative potential term $V_2$ one has a strictly positive potential, which still can be described by eq. (\ref{njl09}), however with a positive coupling $G_{NJL}$. Rewriting that equation as
\begin{eqnarray}
V_{NJL}=-J_A [ (\frac{\bar \psi \psi}{\Lambda_r^3})^2+(\frac{\bar \psi i \gamma_5 \vec \tau \psi}{\Lambda_r^3})^2]
\label{njl229a}
\end{eqnarray}
one should not take the SSB scale $\mu$ as normalization scale any more. Instead another reference scale $\Lambda_r \gg \mu$ has to be introduced which physically corresponds to the distance between two tetrahedrons (alternatively one could utilize the extension of one of them).
One thus obtains an antiferromagnetic internal spin interaction with a negative exchange coupling $J_A$. This repulsion effect leads to the frustrated antiferromagnetic vacuum structure shown in fig. 1. $J_F$ and $J_A$ differ because they correspond to exchange integrals over different regions of space. $J_A$ is dominated by an integral over the volume of one internal tetrahedron (leading to the frustrated internal antiferromagnetic configuration), while $J_F$ is the exchange integral for spin vectors of different tetrahedrons (leading to the aligment of different tehtrahedrons over Minkowski space).
The situation is reminiscent to the theory of ordinary magnets, where the exchange coupling integral J is known to vary with the distance. At large lattice spacings and corresponding large distances between spin vectors ( much larger than the extension of the electron wave function) one has $J>0$ and a ferromagnetic behavior. On the other hand, antiferromagnets like Cr and Mn are characterized by small lattice spacings and corresponding small distances between spin vectors, typically not much larger than the extension of the electron wave function. In these cases $J<0$, i.e. antiferromagnetic behavior.
According to fig. 1 there is one chiral internal spin vector $\vec \pi_i$ for each of the 4 constituents of the internal tetrahedron. In the ground state these vectors point radially away from the origin. Their sum
\begin{equation}
\langle\vec \pi\rangle = \sum_{i=1}^4 \langle\vec \pi_i\rangle
\label{mc7}
\end{equation}
vanishes corresponding to a vanishing vev $\langle\vec \pi \rangle=0$ in accordance with the SM vacuum structure of the Higgs doublet (\ref{hig77}). Excited states arise as vibrations of the vectors $\vec \pi_i$ in fig. 1 and will be interpreted as quarks and leptons. They can be classified according to the system's symmetry group, the Shubnikov group $A_4+S(S_4-A_4)$. This group, which remains unbroken at low energies, has only 1- and 3- dimensional representations, i.e. singlets (interpreted as leptons) and triplets (interpreted as the 3 colors of quarks).
\section{Chiral Extension of the Model}
When studying the dynamics of the internal spin vectors to derive the spectrum of the excited states one notes that $\vec \pi \sim \bar \psi i \gamma_5 \vec \tau \psi$ is not a quantity simple to handle, because it does not fulfill the canonical commutation relations for spin vectors. Secondly, it turns out that the internal Hamiltonian (related to the internal time variable) cannot be written in terms of $\vec \pi$. The appropriate internal vector to use is $\psi^\dagger \gamma_5 \vec \tau \psi$, a well known conserved charge observable from current algebra\cite{current}. However, due to the factor of $\gamma_5$ these vectors still do not fulfill the usual angular momentum commutation relations because their commutator is a scalar, and not a pseudo-scalar any more. In order that the algebra of internal spin vectors closes one is forced to consider the following linear combinations of internal spin vectors
\begin{eqnarray}
\vec Q_R= \frac{1}{\Lambda^3} \psi^\dagger (1+\gamma_5) \vec \tau \psi=a_R^\dagger \vec \tau a_R - b_L^\dagger \vec \tau b_L \\
\vec Q_L=\frac{1}{\Lambda^3} \psi^\dagger (1-\gamma_5) \vec \tau \psi= a_L^\dagger \vec \tau a_L - b_R^\dagger \vec \tau b_R
\label{m1u1}
\end{eqnarray}
where $a(^\dagger)_{L,R}$ and $b(^\dagger)_{L,R}$ denote doublets of annihilation(creation) operators of the fundamental fermion and anti-fermion with left and right handed polarization.
For later use one should discriminate between particle and antiparticle components
\begin{eqnarray}
\vec S&=& a_L^\dagger \vec \tau a_L \qquad \qquad
\vec T= b_L^\dagger \vec \tau b_L \label{njl4baa}\\
\vec U&=& a_R^\dagger \vec \tau a_R \qquad \qquad
\vec V= b_R^\dagger \vec \tau b_R \label{njl4b87}
\end{eqnarray}
These vectors fulfill the canonical algebra of SU(2) generators\cite{current}. The point to note here is the impact of the chiral symmetry group $SU(2)_R\times SU(2)_L$, because in current algebra $\vec Q_L$ and $\vec Q_R$ correspond to the conserved charges of the $SU(2)_R\times SU(2)_L$ symmetry, the factor $\Lambda^{-3}$ arising from the spatial integral of the time component of the left and right handed currents. $\vec S$ and $\vec V$ can be considered as 2 alternative sets of $SU(2)_L$ generators, while $\vec T$ and $\vec U$ serve the same purpose for $SU(2)_R$. For the question of the fermion masses addressed in the next sections one may restrict to consider the vectors $\vec S$ and $\vec T$ because this corresponds to $\bar \psi_R \psi_L$ of the fermion mass term, from which the $\bar \psi_L \psi_R$ part can be obtained by complex conjugation.
Technically, restricting to $\vec S$ and $\vec T$ one is associating the transition between $SU(2)_L$ and $SU(2)_R$ to the charge conjugation operator\cite{nesti}. Alternatively, one could associate it to parity, by restricting to $\vec S$ and $\vec U$. However, this would not lead to the mass terms $\sim \bar \psi_R \psi_L$, and moreover the pairs of Shubnikov states introduced later in eq. (\ref{eq833hg}) could not be interpreted as partners of an $SU(2)_L$ doublet. As well known from the theory of SU(2) representations for a particle operator of the form $a=(U,D)$ the corresponding antiparticle operator is $b=(-D^c,U^c)$, i.e. transitions between particles and antiparticles involve a simultaneous interchange of isospin quantum numbers.
$\vec S$ and $\vec T$ transform under $SU(2)_L$ and $SU(2)_R$, respectively, and fulfill the commutation relations of two decoupled angular momenta in the sense that
\begin{eqnarray}
[S_a,S_b]=i\epsilon_{abc}S_c \qquad
[T_a,T_b]=i\epsilon_{abc}T_c \qquad
[S_a,T_b]=0
\label{sv1}
\end{eqnarray}
On the other hand, within the internal dynamics advocated in this paper $\vec S$ and $\vec T$ play the role of angular momentum observables corresponding to rotations of the internal $R^3$ space. The fact that they are dimensionless according to eqs. (\ref{njl4baa}) agrees with this interpretation, because an angular momentum $\vec r \times \vec p$ is always dimensionless. Physically, the scale $\Lambda$ can be identified as $\Lambda=\mu$ in the SSB regime and $\Lambda_r$ at high energies.
Using eq. (\ref{njl4baa}) one is effectively including further dynamical vibrators $\sim \psi^\dagger \vec \tau \psi$ in addition to the axial vectors $\vec \pi$. This kindly solves another problem not discussed so far, namely the 4$\times$3 d.o.f. of the internal spin vectors in fig. 1 yield only 12 excitation states instead of the necessary 24 quarks and leptons. In order to obtain the remaining 12 (which turn out to be their isospin partners), in ref. \cite{laa} it was proposed that internal displacive vibrations should be included in addition to spin wave excitations. In the present context the doubling of the number of excitations is obtained without displacive vibrations by going to the closed algebra eq. (\ref{mm4xx1}) of the 8 internal spin vectors $\vec S_i$ and $\vec T_i$, i=1-4, whose vacuum values are depicted in fig. 3.
In that figure the vectors $\langle \vec S_i\rangle $ are shown pointing outwards and $\langle \vec T_i\rangle $ pointing inwards fulfilling
\begin{eqnarray}
\langle \vec S_i\rangle = -\langle \vec T_i\rangle
\label{nba77}
\end{eqnarray}
If $\vec S_i$ and $\vec T_i$ were identical observables, this configuration would possess an internal time reversal symmetry (with symmetry group the 'grey' group $S_4 \times \{ 1,S\}$), because the time reversal invariance broken by the set of vectors pointing outwards would be restored by those pointing inwards. However, since $\vec S$ and $\vec T$ are physically different, the ground state has still the original Shubnikov group $A_4+S(S_4-A_4)$ as symmetry.
Eq. (\ref{nba77}) implies that the ground state gets contributions only from the $\gamma_5$ terms in eq. (\ref{njl4baa}) and $\langle \psi^\dagger \vec \tau \psi\rangle_i$ vanishes in the vacuum for each of the constituents $i=1,2,3,4$. In principle the opposite situation is conceivable as well, namely that $\langle \psi^\dagger \vec \tau \psi\rangle_i \neq 0$ while $\langle \psi^\dagger \gamma_5\vec \tau \psi\rangle_i$ vanishes. In that case one would have
\begin{eqnarray}
\langle \vec S_i\rangle = +\langle \vec T_i\rangle
\label{nba77p}
\end{eqnarray}
a perfectly reasonable configuration, which maintains the Shubnikov symmetry for the system of 8 spin vectors as long as all internal vectors point e.g. outwards, in radial directions. In the numerical analysis presented in the following sections the configuration (\ref{nba77p}) will actually be preferred because technically it is easier to handle. In other words, eq. (\ref{nba77p}) will be used as equilibrium conditions for the concrete calculations of masses and eigenstates carried out in the next sections, cf. eq. (\ref{eqq23}).
The 24 eigenmodes of the system can be arranged in six 1-dimensional and six 3-dimensional representations of the Shubnikov group $A_4 + S ( S_4 - A_4)$\cite{shub,borov,white} to yield precisely the multiplet structure of the 24 quark and lepton states of the 3 generations, not less and not more.
\begin{eqnarray}
A_{\uparrow}(\nu_{e})+A_{\uparrow}(\nu_{\mu}) +A_{\uparrow}(\nu_{\tau}) &+&
T_{\uparrow}(u)+T_{\uparrow}(c)+T_{\uparrow}(t)+ \nonumber \\
A_{\downarrow}(e)+A_{\downarrow}(\mu)+A_{\downarrow}(\tau) &+&
T_{\downarrow}(d)+T_{\downarrow}(s)+T_{\downarrow}(b)
\label{eq833hg}
\end{eqnarray}
Here $A_{\uparrow,\downarrow}$ and $T_{\uparrow,\downarrow}$ denote singlet and triplet representations of the Shubnikov group. For simplicity of notation, eq. (\ref{eq833hg}) does not distinguish between particle and anti-particle excitations. As shown later, the $\uparrow$ excitations can be obtained from the $\downarrow$ excitations by interchanging the roles of $\vec S$ and $\vec T$, which according to eq. (\ref{njl4baa}) involves an exchange of particle and anti-particle degrees of freedom. It is well known that for such an exchange the SU(2) quantum numbers have to be exchanged as well, because given an isospin doublet $(u,d)$ the corresponding doublet of anti-particles will be $(-\bar d,\bar u)$. Therefore, apart from describing anti-particles, the $\uparrow$ and $\downarrow$ excitations also describe isospin partners. This is an important feature which carries the isospin quantum number of the fundamental fermion $\psi$ to the quark and lepton isospin wave excitation.
\begin{figure}
\begin{center}
\epsfig{file=gutp24fig3.eps,height=5.4cm}
\caption{The local ground state of the generalized NJL-model eqs.(\ref{n110c}) and (\ref{nba77}). The total of 8 internal spin vectors accounts for 3$\times$8 d.o.f. corresponding to 24 spin vibrations which can be classified according to the multiplet structure of the Shubnikov group, eq.(\ref{eq833hg}). The vectors $\vec S_i^0$ are assumed to point outwards and $\vec T_i^0=-\vec S_i^0$ inwards. According to eq. (\ref{njl4baa}) this corresponds to $\langle \psi^\dagger \gamma_5\vec \tau \psi \rangle_i\neq 0$. The alternative vacuum configuration (\ref{nba77p}) where $\langle \psi^\dagger \vec \tau \psi \rangle_i\neq 0$ and the $\vec T_i^0$ are parallel to the $\vec S_i^0$ instead of anti-parallel is not drawn.}
\nonumber
\end{center}
\end{figure}
In the framework of the chiral NJL dynamics the introduction of the second spin vector corresponds to introducing an additional term including $\bar\psi \vec \tau \psi$ into the potential. Actually it means to consider the most general $SU(2)_L \times SU(2)_R$ invariant potential based on a fundamental isospin doublet $\psi =(U,D)$, the general 2-flavor NJL model\cite{njl2,schechter}
\begin{eqnarray}
V_{2NJL} &=& V_+ +V_- \nonumber \\
V_+&=&-J_+ [ (\frac{\bar \psi \psi}{\Lambda^3})^2+(\frac{\bar \psi \vec \tau \psi}{\Lambda^3})^2
+(\frac{\bar \psi i \gamma_5 \psi}{\Lambda^3})^2+(\frac{\bar \psi i \gamma_5 \vec \tau \psi}{\Lambda^3})^2] \nonumber \\
V_-&=&-J_-[ (\frac{\bar \psi \psi}{\Lambda^3})^2-(\frac{\bar \psi \vec \tau \psi}{\Lambda^3})^2
-(\frac{\bar \psi i \gamma_5 \psi}{\Lambda^3})^2+(\frac{\bar \psi i \gamma_5 \vec \tau \psi}{\Lambda^3})^2]
\label{n110c}
\end{eqnarray}
$V_+$ and $V_-$ are separately chiral $SU(2)_L\times SU(2)_R$ invariant and in addition possess a $U(1)_V$ fermion number symmetry. Furthermore, $V_+$ is invariant under axial $U(1)_A$ transformations, while $V_-$ explicitly breaks this symmetry and can only be used if an axial anomaly is present (see below). The same scale $\Lambda$ ($=\mu$ oder $\Lambda_r$) as in eq. (\ref{njl4baa}) has been introduced to make the fermion operators dimensionless. As before, the NJL couplings can be written in terms of exchange energy densities $J_\pm$, and one needs $J_\pm >0(<0)$ to obtain internal (anti)ferromagnetic behavior.
Rewriting $V_{2NJL}$ as
\begin{eqnarray}
V_{2NJL} =&-&(J_+ +J_-) [ (\frac{\bar \psi \psi}{\Lambda^3})^2
+(\frac{\bar \psi i \gamma_5 \vec \tau \psi}{\Lambda^3})^2] \nonumber \\
&-&(J_+ - J_-)[ (\frac{\bar \psi i \gamma_5 \psi}{\Lambda^3})^2 +
(\frac{\bar \psi \vec \tau \psi}{\Lambda^3})^2]
\label{n110d}
\end{eqnarray}
and using the method of auxiliary fields one can transform the theory in the SSB region into a sigma-model, similar to eq. (\ref{njl09y}), by identifying
\begin{eqnarray}
\sigma &=&-2(G_+ +G_-)\bar \psi \psi \nonumber \\
\vec\pi &=&-2(G_+ +G_-)\bar \psi i \gamma_5 \vec \tau \psi \nonumber \\
\eta &=&-2(G_+ -G_-)\bar \psi i \gamma_5 \psi \nonumber \\
\vec v &=&-2(G_+ -G_-)\bar \psi \vec \tau \psi
\label{njl09w}
\end{eqnarray}
i.e. a scalar iso-scalar field (the physical Higgs $\sigma=\Lambda_F+\phi$), a pseudo-scalar iso-vector (the would be Goldstone bosons $\vec \pi$ absorbed by the weak bosons), a pseudo-scalar iso-scalar $\eta$ and a scalar iso-vector triplet $\vec v$ consisting of 2 charged fields $v^\pm$ and a neutral $v_{z}$.
In other words, the requirement of a closed algebra of interacting spin vectors enforces the introduction of additional scalars which extend the Higgs sector of the SM to a specific two Higgs doublet model\cite{schechter}.
More precisely, the field content indicates a second scalar doublet formed by $\eta$ and $\vec v$ in addition to the ordinary Higgs doublet. However, in order to avoid a chiral ($\gamma_5$) vacuum structure of physical space, the $\eta$-field should not acquire a vacuum expectation value,
so that the second doublet will not be part of the SSB-process, with negative mass terms, a non-trivial minimum of the potential and so on.
Although technically possible, one should not put the octet of fields eq. (\ref{njl09w}) in the adjoint representation
\begin{eqnarray}
\Sigma=\frac{1}{\sqrt{2}}
\begin{pmatrix}
\frac{1}{\sqrt{2}}(\sigma + v_z) & v^+ \\
v^- & \frac{1}{\sqrt{2}}(\sigma - v_z)
\end{pmatrix}
+\frac{i}{\sqrt{2}}
\begin{pmatrix}
\frac{1}{\sqrt{2}}(\eta + \pi_z) & \pi^+ \\
\pi^- & \frac{1}{\sqrt{2}}(\eta - \pi_z)
\end{pmatrix}
\label{higuu}
\end{eqnarray}
of a model with a larger $U(2)_R\times U(2)_L$ symmetry\cite{u2xu2}. The point is that the universal mass term tr$(\Sigma^\dagger \Sigma)$ of such a model would imply $J_-=0$ in eq. (\ref{n110d}). If any, this would be useful only as long as there are no axial anomalies in the theory which break the $U(1)_{A}$ subgroup of $U(2)_R\times U(2)_L$.
As explained in ref \cite{laa} the underlying theory of the present model exhibits such an anomaly.
In general this anomaly not only allows a non-vanishing value for $J_-$ but also makes the mass of the $\eta$ and $\vec v$ much larger than that of the weak gauge bosons and of the Higgs field\cite{u2xu2}. Those bound states may appear as heavy resonances in the TeV regime. They could be interesting dark matter candidates\cite{darkmatter,inert} and play a role at higher energies or higher temperatures of the universe. Their phenomenology, however, will not be discussed at this point, their d.o.f.s just being used as part of the components of the internal spin vectors whose vibrations give the quark and lepton mass spectrum.
\section{Masses and Mixings from Isospin Wave Equations}
The model set up in the last section will now be applied to calculate the quark and lepton masses. The idea is that masses can be identified with eigenfrequencies of excitations of the isospin vectors $\vec S$ and $\vec T$ eqs. (\ref{njl4baa}) and that these eigenfrequencies get contributions both from inner- and from inter-tetrahedral interactions. The {\it inner}-tetrahedral interactions are antiferromagnetic in nature and responsible for the frustrated tetrahedral configuration figs. 1 and 3, i.e. for the structure of the local vacuum. They are small distance contributions and relatively simple to treat because they can be described by an internal antiferromagnetic Heisenberg Hamiltonian for one tetrahedron alone, with corresponding internal spin vector excitations.
On the other hand there are {\it inter}-tetrahedral interactions fed by the 'ferromagnetic' SSB interactions between different tetrahedrons. Their leading effect turns out to be a contribution of order $O(\Lambda_F)$ solely to the top quark mass. Physically speaking, this interaction handicaps the specific eigenmode describing the top quark, because this mode disturbs the SSB alignment in the strongest possible way. Mathematically, the effect will be described by adding an effective universal term to the inner-tetrahedral Heisenberg interaction with a normal ferromagnetic plus a Dzyaloshinskii-Moriya component\cite{dmgut}. The sum of the 2 components will yield a quasi-democratic mass matrix which in leading order only contributes a term of order $\Lambda_F$ to the top-quark mass and nothing to the masses of the other quarks and leptons.
Let me start with the high energy / small distance contributions. The antiferromagnetic inner-tetrahedral Heisenberg Hamiltonian density for the spin vectors $\vec S$ and $\vec T$ reads
\begin{equation}
V_H=- J_{SS}\sum_{i \neq j=1}^4 \vec S_i\vec S_j
- J_{ST}\sum_{i,j=1}^4 [\vec S_i\vec T_j+\vec T_i\vec S_j]
- J_{TT}\sum_{i \neq j=1}^4 \vec T_i\vec T_j
\label{mm3}
\end{equation}
where i and j run over the corners of the tetrahedron fig. 1 and the exchange energy densities J can be identified with the couplings introduced in eq. (\ref{n110d})
\begin{eqnarray}
J_{SS}=J_{TT}=-\frac{1}{2}J_- \qquad\qquad\qquad J_{ST}=\frac{1}{2}J_+
\label{rot22a}
\end{eqnarray}
Using the commutation relation for the internal spin operators
\begin{eqnarray}
[\vec S^a_i,\vec S^b_j]=i\epsilon_{abc}\delta_{ij} S^c_i \qquad
[\vec T^a_i,\vec T^b_j]=i\epsilon_{abc}\delta_{ij} T^c_i \qquad
[\vec S^a_i,\vec T^b_j]=0
\label{mm4xx1}
\end{eqnarray}
one can derive their (internal) time evolution in the Heisenberg picture
\begin{equation}
\Lambda^3 \frac{d\vec S_i}{dt}=i[V_H,\vec S_i] \qquad\qquad\qquad
\Lambda^3 \frac{d\vec T_i}{dt}=i[V_H,\vec T_i]
\label{mm42}
\end{equation}
to obtain
\begin{eqnarray}
\frac{d\vec S_i}{dt}= \vec S_i^0 \times \sum_{i \neq j=1}^4 [j_{SS}\vec S_j+j_{ST}\vec T_j]+k_{ST}\vec S_i^0\times \vec T_i \label{eqq22vv} \\
\frac{d\vec T_i}{dt}= \vec T_i^0 \times \sum_{i \neq j=1}^4 [j_{TT}\vec T_j+j_{ST}\vec S_j]+k_{ST}\vec T_i^0\times \vec S_i
\label{eqq22}
\end{eqnarray}
These equations have been linearized for small displacements $\delta \vec S_i=\vec S_i-\vec S_i^0$ and $\delta \vec T_i=\vec T_i-\vec T_i^0$ of the spin vectors from their ground state positions $\vec S_i^0 =\langle \vec S_i\rangle$ and $\vec T_i^0 =\langle \vec T_i\rangle$ in fig. 3, and the letter $\delta$ has then been left out. Furthermore, we have switched from exchange energy densities
$J$ to exchange energies
\begin{eqnarray}
j=J\Lambda^{-3}
\end{eqnarray}
Finally, eq. (\ref{eqq22}) includes a contribution, which accounts for the possibility that $\vec S_i \vec T_i$ interact with a different strength than $\vec S_i \vec T_j$ for $j\neq i$, because the internal distance between $\vec S_i$ and $\vec T_i$ is different from the distance between $\vec S_i$ and $\vec T_j$ for $j\neq i$.
The ground state positions are given by
\begin{eqnarray}
\vec S_1^0=\frac{1}{\sqrt{3}}(-1,-1,-1) \qquad\vec S_2^0=\frac{1}{\sqrt{3}}(-1,+1,+1) \\
\vec S_3^0=\frac{1}{\sqrt{3}}(+1,-1,+1) \qquad\vec S_4^0=\frac{1}{\sqrt{3}}(+1,+1,-1)
\label{eqq23}
\end{eqnarray}
and $\vec T_i^0=\pm\vec S_i^0$ depending on whether one is analyzing the parallel or anti-parallel configuration fig. 3. An overall normalization factor N of the vacuum spin vectors was put to 1, because it does not influence the eigenfrequencies and mixing angles. More precisely, it can be absorbed by a redefinition of the couplings $j\rightarrow jN$.
So everything fixed now to solve the differential equations (\ref{eqq22})? Not quite, because for a frustrated, i.e. non-collinear ground state configuration like fig. 1 it is necessary to go beyond the collinear spin wave analysis and transform to a rotating frame with the z-axis pointing along the local spin direction\cite{dmz1,dmz2}. Applied to the present case this procedure modifies eqs. (\ref{eqq22}) in such a way that
\begin{eqnarray}
\frac{d\overrightarrow{U_iS_i}}{dt}= \overrightarrow{U_iS_i} \times \sum_{i \neq j=1}^4 [j_{SS}\overrightarrow{U_jS_j}+j_{ST}\overrightarrow{U_jT_j}] +k_{ST}\overrightarrow{U_iS_i}\times \overrightarrow{U_iT_i} \label{eqq24vv} \\
\frac{d\overrightarrow{U_iT_i}}{dt}= \overrightarrow{U_iT_i} \times \sum_{i \neq j=1}^4 [j_{TT}\overrightarrow{U_jT_j}+j_{ST}\overrightarrow{U_jS_j}] +k_{ST}\overrightarrow{U_iT_i}\times \overrightarrow{U_iS_i}
\label{eqq24}
\end{eqnarray}
where $U_i$ are diagonal $3\times 3$ matrices which act on the vector components of the spin vectors
\begin{eqnarray}
U_1=D(1,1,1) \quad U_2=D(1,-1,-1) \quad
U_3=D(-1,1,-1) \quad U_4=D(-1,-1,1)
\label{eqq25}
\end{eqnarray}
The resulting 24$\times$24 matrix has to be diagonalized in order to obtain the 24 eigenfrequencies $\omega$. Due to the Shubnikov symmetry of the system the corresponding eigenstates can be arranged into 6 singlets and 6 triplets as in eq. (\ref{eq833hg}), i.e. as leptons and quarks. Each triplet consists of 3 states with degenerate eigenvalues, because the Shubnikov symmetry $A_4+S(S_4-A_4)$ is unbroken.
The result of the diagonalization procedure gives the following non-vanishing masses / eigenfrequencies consisting of 2 singlets
\begin{eqnarray}
\omega (\mu) = -\omega (\tau)&=& \pm(6 j_{ST}+ 2 k_{ST})
\label{mm15f}
\end{eqnarray}
and 4 triplets
\begin{eqnarray}
\omega (c) = -\omega (t) &=& \sqrt{j_1^2 + \Delta j_2 }
\label{mm15ff44}
\\
\omega (s) = -\omega (b) &=& \sqrt{j_1^2 - \Delta j_2 }
\label{mm15ff}
\end{eqnarray}
where $j_1^2$, $\Delta$ and $j_2$ are abbreviations for
\begin{eqnarray}
j_1^2&=&-8(j_{SS}^2+j_{TT}^2)-10j_{ST}^2-2k_{ST}^2-2(j_{SS}+j_{TT})(6j_{ST}+2k_{ST})
-2j_{ST}k_{ST} \nonumber \\
\Delta^2&=&4(j_{SS}-j_{TT})^2+(j_{ST}-k_{ST})^2 \nonumber \\
j_2&=&4(j_{SS}+j_{TT})+6 j_{ST}+ 2 k_{ST}
\end{eqnarray}
The corresponding eigenvectors are not given because the formulas are too cumbersome to be presented here.
At this stage, apart from these 14 modes there are 10 zero modes (4 singlets and 2 triplets), which can be attributed to the 3 neutrinos, the electron and the up- and down-quark.
Using $j_{SS}=j_{TT}$ obtained in eq. (\ref{rot22a}) from the heuristic deduction via the generalized NJL model the quark mass eigenvalues simplify to
\begin{eqnarray}
\omega (c) = -\omega (t) &=&4j_{SS} +4j_{ST}
\label{mm15ff44g}
\\
\omega (s) = -\omega (b) &=&4j_{SS}+2 j_{ST} +2k_{ST}
\label{mm15ffg}
\end{eqnarray}
One concludes that considering only the anti-ferromagnetic Heisenberg contributions eqs. (\ref{mm3}) and (\ref{eqq22}) leads to a stronger degeneracy than dictated by the Shubnikov symmetry alone. In fact, this does not only concern the zero modes, because according to (\ref{mm15f})-(\ref{mm15ff}) the quarks and leptons of the second and third family have the same mass.
In the next section effects of inter-tetrahedral interactions will be included to partially lift this degeneracy between the families and in particular to shift the masses of the third family to larger values. Most prominently, the top quark mass will be equipped with a contribution of order $\Lambda_F$. Afterwards, torsion and anisotropic corrections will be included. They are tiny effects and cannot be attributed to a SM piece of the interactions like the contributions discussed so far. However, they are needed, because they are responsible for the light quark and lepton masses, and in particular for the neutrino masses and mixings. Hence they will remove the 'accidental' degeneracies which one obtains if one only considers the inner-tetrahedral Heisenberg exchange contributions to the eigenfrequencies.
What can be done already at this point is to take the formulas (\ref{mm15f})-(\ref{mm15ff}) for the second family and to adopt them to the known mass values of the muon, the charmed and the strange quark. It is advisable to take the running values of the masses at order 1 TeV\cite{runningmass} and try to determine from that the values for the internal exchange couplings. A closer look at eqs. (\ref{mm15f})-(\ref{mm15ff}) reveals that there is no simple solution with small or antiferromagnetic (i.e. negative) value of $k_{ST}$. Instead, one obtains
\begin{eqnarray}
j_{ST}\approx -0.12 \ \text{GeV} \qquad j_{TT}=j_{SS}\approx -0.12 \ \text{GeV} \qquad k_{ST} \approx 0.32 \ \text{GeV}
\label{mm15ffff}
\end{eqnarray}
and concludes that the coupling strengths are $\leq 1$ GeV with a ferromagnetic coupling $k_{ST}$ of adjacent spin vectors $\vec S_i$ and $\vec T_i$, while the interactions with $i\neq j$ are somewhat smaller in magnitude and anti-ferromagnetic. This is precisely, what was anticipated for the vacuum configuration eq. (\ref{nba77p}).
\section{Top Quark Mass from Dzyaloshinskii Moriya Interactions}
As discussed before, the {\it inter}-tetrahedral interactions should yield the leading SSB contributions to the fermion masses. To account for these contributions one has to include a sum over neighboring tetrahedrons in the spin Hamiltonian and add the effect of the corresponding interactions to the e.o.m. Since I do not have the resources to treat these many terms properly the following trick will be used to approximately solve the problem: the effects of all other tetrahedrons on a given tetrahedron fig. 3 will be subsumed as an effective contribution generated by the internal spins. The idea is that this effective contribution can be attributed to the gauge transformation eq. (\ref{hipoga}) which transfers the $\vec\pi$-field from the Higgs sector to the W-boson mass term. As well known such a transformation modifies the W-field by
\begin{eqnarray}
\vec W_\mu \rightarrow \vec W_\mu - \frac{1}{g\Lambda_F}\partial_\mu \vec\pi - \frac{1}{\Lambda_F}\vec\pi \times \vec W_\mu
\label{mm19u1}
\end{eqnarray}
Thus, while the bilinear terms in $\vec\pi$ disappear from the Higgs potential eq. (\ref{hipo}), they re-appear in the W-mass term of the Lagrangian. Furthermore, their sign is such that the antiferromagnetic $\vec\pi - \vec\pi$ coupling at high energies gets transformed into a ferromagnetic interaction plus an additional term which is due to the non-abelian nature of the gauge transformation. All in all
\begin{equation}
\frac{1}{2} m_W^2\vec W_\mu \vec W^\mu \rightarrow \frac{\Lambda_F}{8} \sum_{i,j=1}^4 [ \vec S_i\vec S_j +i (\vec S_i\times \vec D_{ij}) \vec S_j +c.j. ]
\label{mm3dm}
\end{equation}
has to be added to the Heisenberg Hamiltonian (\ref{mm3}). To obtain this result the tree level relation $g=2m_W/\Lambda_F$ has been used. Note that because of the $SU(2)_L$ interactions only terms involving $\vec S$ and its charge conjugate (c.j.) $\vec V$ appear, as defined in (\ref{njl4b87}), in accordance with the V-A structure of the weak interactions.
This is in agreement with ref. \cite{laa} where it has been shown that after the formation of the chiral tetrahedrons the internal spin $SU(2)$ becomes a symmetry involving only left handed particles, $SU(2)_L$.
The additional term in eq. (\ref{mm3dm}) involving the cross product can be interpreted as a Dzyaloshinskii-Moriya component\cite{dmgut}, a contribution which in solid state physics is sometimes used to describe leading anisotropic corrections to the ordinary Heisenberg equations of motion. Quite in general, such a component stands for a tendency to form a rotational structure (instead of the ordinary ferromagnetic alignment of neighboring tetrahedrons depicted in fig. 2) simply because the DM-term tends to rotate the spin vectors instead of aligning them. In the present framework it was deduced as a consequence of the gauge transformation eq. (\ref{hipoga}). Therefore the DM-term can be interpreted quite naturally, namely by that the $SU(2)_L$ gauge fields induce a curvature of the fiber bundle formed by the system of all tetrahedrons, and the DM-term simply takes care of this curvature effect to effectively maintain the aligned structure.
In general, a DM-component to the Heisenberg Hamiltonian has a complicated coupling structure involving vector couplings $\vec D_{ij}$, as given in eq. (\ref{mm3dm}). However, in the case at hand the couplings $\vec D_{ij}$ are fixed by 2 symmetry requirements, namely that the term must be $SU(2)_L$ gauge invariant and that it must respect the $A_4+S(S_4-A_4)$ Shubnikov symmetry. While the former fixes the modulus of the DM coupling strength relative to the Heisenberg term in eq. (\ref{mm3dm}), the latter forces the direction of $\vec D_{ij}$ to be $\vec S^0_k-\vec S^0_l$\cite{dmgut}, where $kl\neq ij$ is chosen such that the sign of the permutation (ijkl) of (1234) is positive\cite{dmgut}.
The equations of motion of the spin vectors (\ref{eqq22vv}) are supplemented by the inter-tetrahedral contribution (\ref{mm3dm}) in the following way:
\begin{eqnarray}
\frac{d\vec S_i}{dt}= \frac{1}{4} \Lambda_F \{ \vec S_i \times ( [ 1+ i \sum_{j}\vec D_{ij} \times ] \vec S_j ) \}
\label{eqq212}
\end{eqnarray}
These equations modify only the 12 equations (\ref{eqq22vv}) for $\vec S_i$ leaving the equations for the $SU(2)_R$ vibrators $\vec T_i$ (\ref{eqq22}) untouched. Their net effect is to give a quasi-democratic type of contribution $\sim \Lambda_F$ to the mass matrix, in the sense that each entry of the 12$\times$12 eigenmatrix for $\vec S_i$ gets a contribution involving $\Lambda_F$. Evaluating the eigenvalues, this equips the top quark triplet with a mass
\begin{eqnarray}
m_t=N \Lambda_F=\frac{\Lambda_F}{2g}=\frac{\Lambda_F^2}{4m_W} \approx 190 GeV
\label{eqqmt}
\end{eqnarray}
while leaving the other quark and lepton masses unchanged.
An overall factor $N$ has been included which arises from the normalization of the spin vectors and was left out in eqs. (\ref{eqq23}) and (\ref{mm15f})-(\ref{mm15ffg}) because it was absorbed in the redefinition of the Heisenberg couplings. This is not possible in the case at hand, because the coupling is given in terms of the predefined quantity $\Lambda_F$. $N$ can be shown to be related to the $SU(2)_L$ gauge coupling g via $2gN=1$.
Now that we have established a top quark mass of order $\Lambda_F$ one may ask whether there are contributions to the masses of the $\tau$-lepton and the b-quark from the inter-tetrahedral couplings as well. Such contributions are strongly desired, because one would like to get rid of the relations $m_b=m_s$ and $m_\tau=m_\mu$ following from eqs. (\ref{mm15f}) and (\ref{mm15ff}). Unfortunately, if one sticks to (\ref{mm3dm}) and (\ref{eqq212}), the answer to that question is no. These equations only equip the top quark with a mass.
However, one should keep in mind that (\ref{mm3dm}) and (\ref{eqq212}) represent a rather crude approximation to the inter-tetrahedral interaction effects. One may analyze several possible modifications on whether they lead to the desired effect. For example, one may think of a contribution due to $\gamma$-Z mixing. As well known, this mixing leads to a modification of the W-mass term in the form $\sim m_Z^2 Z_\mu Z^\mu$ where $m_Z:=m_W/\cos \theta_W$ and $Z:=W_z\cos\theta_W-B\sin\theta_W$ and $B$ is the SM $U(1)_Y$ gauge field. One obviously has $m_Z^2 Z_\mu Z^\mu=m_W^2 W_{z\mu} W_z^\mu+O(B)$. In effect the spin-spin interaction eq. (\ref{mm3dm}) is not modified, the physical reason being that the $U(1)_Y$ terms are isospin blind and therefore do not modify directions in the internal space.
Another possibility are mixings among the various fields in eq. (\ref{njl09w}). In fact, the quantum numbers of the Higgs allow for a mixing with the z-component of the iso-vector $\vec v$. Similarly, the pseudo-scalar combination $\eta\sim \bar\psi i\gamma_5 \psi$ can mix with $\pi_z$. However, according to my analysis there is again no effect on the quark and lepton masses.
Finally there may be an admixture of a right handed current in such a way that $\vec S$ in eq. (\ref{mm3dm}) should effectively be replaced by
\begin{eqnarray}
\vec S \rightarrow \vec S+\alpha \vec T
\label{eqqrh}
\end{eqnarray}
It is true that there are rather strong restrictions to the presence of right handed currents and on the value of $\alpha$, but effects in the percent range are still allowed \cite{nesti,moha2}.
I am not claiming that they are actually present, in particular, because a right-handed W-boson contradicts the arguments advocated in ref. \cite{laa}, that the handedness of the tetrahedral structure fig. 1 should make any weak interaction left-handed. I am just considering them for the sake of having a definite example, because such terms turn out to have the nice feature that the mass contribution to the top quark essentially remains unchanged while new contributions to the b- and $\tau$-mass are generated of order $\sim\alpha\Lambda_F$. More precisely, the contribution from this source is
\begin{eqnarray}
m_b=m_\tau= \frac{3\alpha\Lambda_F^2}{16m_W}
\label{eqqrh6}
\end{eqnarray}
which interestingly is in accord with the Georgi-Jarlskog mass relations\cite{georgi}. Note that all lighter quark and lepton masses are not modified by the $\alpha\vec T$ term in eq. (\ref{eqqrh}).
\section{Neutrino Masses and the Conservation of internal Angular Momentum}
The calculations presented so far are based on a certain interpretation of the Standard Model Higgs mechanism - a rather intuitive interpretation, if one is willing to accept that the form of the quark and lepton spectrum is due to a discrete internal tetrahedral structure. As was shown above it is possible to identify the terms in the SM Lagrangian responsible for the internal spin interactions, namely the quadratic part of the Higgs potential and the W-mass term. However, at this point 10 of the 24 possible excitations (the neutrinos, the electron and up and down quarks) are still without a mass.
As long as these 4 singlets and 2 triplets remain massless, i.e. constant, non-vibrating modes, it is no use trying to calculate the CKM matrix elements or even the mixing angles in the neutrino sector. To get rid of those degeneracies one has to relax a condition inherent in the classical Heisenberg model namely that the internal magnetic moments may be treated as classical 3-dimensional vector spins of fixed length. This condition is destroyed by quantum fluctuations in the quantum Heisenberg model and on the classical level by allowing for torsional vibrations.
Although these (tiny) torsional effects have no counterpart in the SM Lagrangian it can be shown that the leading up-quark, down-quark and electron mass contributions are provided by isotropic torsional interactions of the internal spin vectors while the differences in the neutrino masses can be attributed to anisotropic effects within the torsional couplings.
To start with I am now going to write down the most general form of these torsional interactions.
As argued above, torsion is not strictly forbidden in the system under consideration, and for the case of only one spin vector is simply induced by a contribution of the form $d\vec S /dt \sim \vec S$ to its time variation. In the case at hand with 8 spin vectors $\vec S_i$ and $\vec T_i$ its main effect is to allow vibrations along the local z-directions and thus lift the degeneracies of all zero modes. The terms supplementing the Heisenberg e.o.m. (\ref{eqq22vv}) and (\ref{eqq22}) are
\begin{eqnarray}
\frac{d\vec S_i}{dt}= i e_{SS} \vec S_i +if_{SS} \sum_{j\neq i} \vec S_j+ie_{ST} \vec T_i +if_{ST} \sum_{j\neq i} \vec T_j \label{eqq22ii} \\
\frac{d\vec T_i}{dt}= i e_{TT} \vec T_i +if_{TT} \sum_{j\neq i} \vec T_j+ie_{ST} \vec S_i +if_{ST} \sum_{j\neq i} \vec S_j
\label{eqq22iii}
\end{eqnarray}
where e and f are the torsion coupling strengths, whose values are assumed to be small compared to the exchange couplings considered so far. More precisely one has the natural hierarchy
\begin{eqnarray}
e,f \sim O(MeV) \ll j,k \sim O(GeV) \ll \Lambda_F \alpha \ll \Lambda_F
\label{eqq2113}
\end{eqnarray}
so that the torsional couplings can indeed be expected to provide for the electron and the up- and down-quark mass with $m_{e,u,d}/m_{\mu,s,c} \sim 10^{-2}$. As will be shown in the next section neutrino masses are due to still smaller anisotropy effects among the torsional couplings.
The contributions (\ref{eqq22ii}) and (\ref{eqq22iii}) can be incorporated in the full 24$\times$24 eigenvalue problem to yield
\begin{eqnarray}
m_e &=& e_{SS}-e_{ST}+3f_{SS}-3f_{ST} \label{mas86} \\
m_\mu &=& 6 j_{ST}+ 2 k_{ST}+O(m_e) \\
m_\tau &=& \frac{3\alpha\Lambda_F^2}{16m_W} + O(m_\mu)
\end{eqnarray}
\begin{eqnarray}
m_u &=& e_{SS}+e_{ST}-f_{SS}-f_{ST} \\
m_c &=& 4j_{SS} +4j_{ST} +O(m_u) \\
m_t &=& \frac{\Lambda_F^2}{4m_W} + O(m_c)
\end{eqnarray}
\begin{eqnarray}
m_d &=& e_{SS}-e_{ST}-f_{SS}+f_{ST} \\
m_s &=& 4j_{SS}+2 j_{ST} +2k_{ST}+O(m_d) \\
m_b &=& \frac{3\alpha\Lambda_F^2}{16m_W} + O(m_s)
\end{eqnarray}
\begin{eqnarray}
m(\nu_e) &=& m(\nu_\mu) = m(\nu_\tau) =e_{SS}+e_{ST}+3f_{SS}+3f_{ST}
\label{mas8}
\end{eqnarray}
For simpicity and to be in accord with the NJL expectation $j_{SS}=j_{TT}$ eq. (\ref{rot22a}) I have set $e_{SS}=e_{TT}$ and $f_{SS}=f_{TT}$.
It is interesting to note that eq. (\ref{mas8}) seems to indicate that the neutrinos acquire a mass, too - in fact the same mass for all neutrino species. Actually, the quantity on the right hand side of (\ref{mas8}) deserves special attention, because it governs the evolution of the total internal angular momentum $\vec \Sigma:=\sum_i (\vec S_i+\vec T_i)$ as can be seen by adding up all contributions eqs. (\ref{eqq22ii}) and (\ref{eqq22iii})
\begin{eqnarray}
\frac{d \vec \Sigma}{dt}\equiv \frac{d\sum_{i=1}^4 (\vec S_i+\vec T_i)}{dt}= i (e_{SS}+e_{ST}+3f_{SS}+3f_{ST}) \sum_{i=1}^4 (\vec S_i+\vec T_i)
\label{eqsu8}
\end{eqnarray}
It should be noted that the Heisenberg interactions (\ref{mm3}) as well as the DM term (\ref{eqq212}) fulfill $d\vec\Sigma /dt=0$ and thus do not contribute to eq. (\ref{eqsu8}).
Comparing (\ref{mas8}) and (\ref{eqsu8}) one concludes, that the 3 neutrino eigenstates correspond to the vibrations of the 3 components of $\vec \Sigma$. Whenever the total internal angular momentum is conserved, i.e. $\vec \Sigma$ is independent of the (internal) time, the neutrino masses will strictly vanish, while non-zero neutrino masses are a signal for non-conservation of $\vec \Sigma$.
In the framework of Noether's Theorem $\vec \Sigma$ is the conserved charge related to the internal rotational symmetry, and the neutrino states $\nu_e$, $\nu_\mu$ and $\nu_\tau$ are related to the breaking of this continuous symmetry. Applying Goldstone's theorem to the internal dynamics yields 3 internal massless magnon excitations, in a similar way, as magnons are obtained as Goldstone bosons of the broken rotational symmetry in ordinary magnetic systems.
Goldstone bosons? This sounds strange in view of the fact that neutrinos are fermions. The point is that one has to distinguish the dynamics in internal from that in physical space. In physical space the neutrinos are fermions, but in internal space they are described by (bosonic) excitations of the internal angular momentum $\vec S +\vec T$ which is the conserved quantity associated with the internal rotational symmetry.
In principle, the general solution to the eigenproblem does not only give the eigenvalues (\ref{mas86})-(\ref{mas8}) but via the corresponding eigenvectors can also be used to accommodate all physical quark and lepton mixing parameters. Due to lack of resources I have to leave the determination of the quark mixings
to future work. To get some understanding of the physics of the mixing I will now concentrate on the lepton sector of the theory, which is somewhat easier to handle than the full 24$\times$24 problem. To obtain non-vanishing differences between the masses of the neutrino species additional interaction terms violating internal angular momentum conservation have to be introduced. Afterwards, the method how to determine the PMNS mixing matrices from these interactions within the present framework will be desribed in detail.
\section{(Not just) a Toy Model for Leptons}
The mass problem for the leptons alone can be approximately reduced from the tetrahedral configuration to the simple 1-dimensional structure depicted in fig. 4. As Eq. (\ref{mm15f}) indicates, the lepton masses do not depend on the couplings $J_{SS}$ and $J_{TT}$ but only on $J_{ST}$.
As a matter of fact using some simple matrix algebra manipulations it may be shown that the Heisenberg contributions to the Shubnikov singlets (leptons) can effectively be obtained from the configuration of fig. 4 with only 2 internal spin vectors $\vec S$ and $\vec T$ (instead of 8) which in the ground state point in the z-directions
\begin{equation}
\langle\vec S\rangle=\langle\vec T\rangle=(0,0,1)
\label{mm5}
\end{equation}
To analyze the behavior of this system, I will start with the Heisenberg part of the interactions
\begin{equation}
V_H=-J \vec S \vec T
\label{mm655}
\end{equation}
where $J$ is identical to $4J_{ST}$ used in the last section. The factor of $4$ is a geometrical factor arising from the reduction of the tetrahedral configuration.
\begin{figure}
\begin{center}
\epsfig{file=gutp24fig444.eps,height=3.5cm}
\caption{The local ground state in the model for leptons, where the spin vectors $\vec S$ and $\vec T$ point in the z-direction. The parallel configuration is depicted here instead of the anti-parallel one in fig. 3. If only a Heisenberg interaction of the form (\ref{mm655}) is included the combination $\vec S +\vec T$ remains static, thus giving rise to 3 zero modes corresponding to the 3 neutrinos. A fourth zero-mode appears in this limit, because the spin vectors will rotate only in the transversal $(x,y)$-plane, leaving all torsional vibrations (i.e. in the z-direction) as zero modes. If further interactions are added to the Heisenberg term, all 6 eigenmodes receive non-vanishing values.}
\nonumber
\end{center}
\end{figure}
The 6 d.o.f. of this system of 2 internal chiral spin vectors lead to 6 eigenmodes, and we are now going to show how the lepton masses and mixings, in particular the tiny neutrino masses may arise. The symmetry group of the ground state fig. 4 is $\{1,SR\}$, whose only non-trivial element is a reflection R at the x-y-plane followed by an internal time reversal S.
This group has only singlet representations\cite{shub}, according to which the 6 eigenmodes will be classified.
\begin{table}
\begin{center}
\begin{tabular}{|l|c|c|c|c|c|c|}
\hline
&$S_x$&$S_y$&$S_z$&$T_x$&$T_y$&$T_z$\\
\hline
$S_x$ & $i(\omega_f+v)$ & $j+v$ & $0$ & $i(f-v)$ & $-j-v$ & $0$ \\
$S_y$ & $-j-v$ & $i(\omega_f+v)$ & $0$ & $j+v$ & $i(f-v)$ & $0$ \\
$S_z$ & $0$ & $0$ & $i\omega_f$ & $0$ & $0$ & $if$ \\
$T_x$ & $i(f-v)$ & $-j-v$ & $0$ & $i(\omega_f+v)$ & $j+v$ & $0$ \\
$T_y$ & $j+v$ & $i(f+v)$ & $0$ & $-j-v$ & $i(\omega_f+v)$ & $0$ \\
$T_z$ & $0$ & $0$ & $if$ & $0$ & $0$ & $i\omega_f$ \\
\hline
\end{tabular}
\bigskip
\caption{The interaction matrix between the 2 chiral spin vectors $\vec S$ and $\vec T$ of figure 4 giving rise to electron-, muon- and tau-mass. In addition to the Heisenberg interaction eq. (\ref{mm655}) inter-tetrahedral effects ($v$) and a universal torsional coupling ($f$) have been introduced. The neutrinos are still massless at this stage, corresponding to the 3 d.o.f. of $\vec S+\vec T$ which do not vibrate. The abbreviation $\omega_f=\omega -f$ is used.}
\label{tab331}
\end{center}
\end{table}
Adding a small universal torsional coupling $f \ll j$ the time evolution of the spin vectors is given by
\begin{equation}
\frac{d\vec S}{dt}=j \vec S \times \vec T +f (\vec S -\vec T)\qquad\qquad
\frac{d\vec T}{dt}=j \vec T \times \vec S+f (\vec T-\vec S)
\label{mm7nor}
\end{equation}
One then has to diagonalize the sum of the matrices given in table \ref{tab331} in order to obtain the eigenstates
\begin{eqnarray}
\nu_e&=&(0, 0, 1, 0, 0, 1) \qquad
e=(0, 0, -1, 0, 0, 1) \nonumber \\
\nu_\mu&=&(0, 1, 0, 0, 1, 0) \qquad
\mu=(-i, -1, 0, i, 1, 0) \nonumber \\
\nu_\tau&=&(1, 0, 0, 1, 0, 0) \qquad
\tau=(i, -1, 0, -i, 1, 0)
\label{mm23rr}
\end{eqnarray}
and their masses/eigenfrequencies
\begin{eqnarray}
\omega(\nu_e)&=&0 \qquad\qquad\qquad
\omega(e)=2f \nonumber \\
\omega(\nu_\mu)&=&0 \qquad \qquad\qquad
\omega(\mu)=2(j+f) \nonumber \\
\omega(\nu_\tau)&=&0 \qquad \qquad\qquad
\omega(\tau)= 2(2v-j-f)
\label{eqq2260}
\end{eqnarray}
In eq. (\ref{mm23rr}) a 6-dimensional vector space of eigenvectors has been introduced in which the sum and difference $\vec S \pm \vec T$ are simply given by
\begin{eqnarray}
S_x\pm T_x=\frac{1}{\sqrt{2}}(1,0,0,\pm 1,0,0) \nonumber \\
S_y\pm T_y=\frac{1}{\sqrt{2}}(0,1,0,0,\pm 1,0) \nonumber \\
S_z\pm T_z=\frac{1}{\sqrt{2}}(0,0,1,0,0,\pm 1)
\label{mm23h}
\end{eqnarray}
\begin{table}
\begin{center}
\begin{tabular}{|l|c|c|c|c|c|c|}
\hline
&$S_x$&$S_y$&$S_z$&$T_x$&$T_y$&$T_z$\\
\hline
$S_x$ & $i e_{xy}$ & $-j_z$& $j_{xy}$ & $if_{xy}$ & $-l_z$& $-l_{xy}$ \\
$S_y$ & $j_z$ & $ie_{xy}$ & $-j_{xy}$ & $l_z$ & $if_{xy}$&$l_{xy}$ \\
$S_z$ & $-j_{xy}$ &$j_{xy}$ & $ie_z$ & $l_{xy}$ & $-l_{xy}$& $if_z$ \\
$T_x$ & $if_{xy}$ &$-l_z$& $-l_{xy}$& $ig_{xy}$ & $-m_z$& $m_{xy}$ \\
$T_y$ & $l_z$&$if_{xy}$& $l_{xy}$ & $m_z$& $ig_{xy}$ & $-m_{xy}$ \\
$T_z$ & $l_{xy}$& $-l_{xy}$& $if_z$ & $-m_{xy}$& $m_{xy}$& $ig_z$ \\
\hline
\end{tabular}
\bigskip
\caption{The most general correction terms to the matrix in table \ref{tab331}
}
\label{tab3322}
\end{center}
\end{table}
Furthermore, a contribution $v\sim \alpha \Lambda_F$ reflecting the leading inter-tetrahedral contribution to the $\tau$ mass has been included in table \ref{tab331} and eq. (\ref{eqq2260}) which lifts the degeneracy between muon and $\tau$-lepton. The natural hierarchy between these couplings can then be invoked to accommodate the lepton masses. The electron mass is naturally small as compared to $m_\mu$ and $m_\tau$ because the electron corresponds to a torsional vibration (of $S_z - T_z$) in the z-direction and its mass gets only torsional contributions $\sim f \approx 0.25$MeV. There is no contribution $\sim j$ to the electron mass because the Heisenberg interaction conserves each spin's fixed length and does not allow spin vibrations in the z-direction. Note the other mode in the z-direction, the one $\sim S_z + T_z$, is to describe the electron neutrino.
The Hamiltonian on which eq. (\ref{mm7nor}) is based conserves the total spin of the system so that the 3 d.o.f. corresponding to this quantity do not vibrate. Comparing (\ref{mm23rr}) with (\ref{mm23h}) one explicitly sees that the neutrino zero modes correspond to the 3 components of the conserved total internal angular momentum $\vec S +\vec T$. As discussed in detail in the last section they can be interpreted as Goldstone modes of the broken internal rotational invariance. In order to obtain non-zero neutrino masses one may add a small contribution $\sim \vec S +\vec T$ to the right hand side of eq. (\ref{mm7nor}). Similar to (\ref{mas8}) this would equip all neutrino species with the same mass.
In order to be more general and obtain different masses for the 3 neutrinos I have written down the most general interaction matrix which violates internal angular momentum conservation and includes anisotropic and torsional forces in table \ref{tab3322}. Compared to the leading terms in table \ref{tab331} the new contributions must be tiny, as are the neutrino masses. More precisely, they should be of order at most $ O(m_\nu /m_e)\approx 10^{-7}$.
It seems clear that corrections of such minuteness are difficult to handle quantitatively. Nevertheless, the present approach allows to analyze the question from which of the various sources appearing in table \ref{tab3322} the observed neutrino masses and mixings\cite{nuemix} actually arise. For example, an inverted hierarchy of neutrino masses which seems to be slightly favored by the present data can be accommodated quite easily, and measured values of the mixing angles will further determine the coupling parameters in table \ref{tab3322}.
A complete numerical analysis of the whole parameter space is not undertaken in the present study. Rather, one realistic example will be discussed to see whether the neutrino measured parameters can be reproduced. Namely, a simple solution to the case of the inverted hierarchy is obtained by putting $j_{xy}=l_{xy}=m_{xy}\equiv \delta_j$ and $f_{xy}\equiv \delta_e$ and all other parameters in table \ref{tab3322} to zero. The result for the masses then is
\begin{eqnarray}
\omega(\nu_e)&=&-\delta_e \qquad \qquad\qquad\qquad \qquad\qquad
\omega(e)=2f \\
\omega(\nu_\mu)&=&-\frac{1}{2}\{\delta_e+\sqrt{\delta_e^2+32\delta_j^2}\} \qquad \qquad
\omega(\mu)=2(j+f)+\delta_e \nonumber \\
\omega(\nu_\tau)&=& -\frac{1}{2}\{\delta_e-\sqrt{\delta_e^2+32\delta_j^2}\} \qquad\qquad
\omega(\tau)= 2(2v+j-f)+\delta_e \nonumber
\label{mm15}
\end{eqnarray}
Assuming $\delta_e \gg \delta_j$ one can indeed reproduce the 'inverted hierarchy' of neutrino masses. To be definite we choose the values $\delta_j=0.002$eV, $\delta_e=0.05$eV, $f=0.25$MeV $j=50$MeV $v=0.4$GeV in order to reproduce the current data, i.e.
\begin{eqnarray}
m_2^2-m_1^2=0.000063\ eV^2 \\
m_3^2-\frac{1}{2}(m_1^2+m_2^2)=-0.0025\ eV^2
\label{mm1580}
\end{eqnarray}
Next one has to check whether the mixing angles for the neutrino come out right. This is then the appropriate place to discuss the general strategy how mixing matrices can be obtained in the present framework. To determine the CKM quark mixing elements the eigenvectors of a complicated 24$\times$24 problem have to be fixed. For the case of leptons with the 6$\times$6 matrix in tables \ref{tab331} and \ref{tab3322} the calculation of eigenvectors and PMNS mixing matrix elements is a relatively simple exercise.
One just has to remember that the mixing matrix is defined as the unitary transition matrix $U_N=(U_{N\alpha a})$ between the mass eigenstates $\nu_\alpha, \alpha=e,\mu,\tau$, and the weak interaction eigenstates $\nu_a, a=1,2,3$:
\begin{eqnarray}
\nu_\alpha = \sum_{a} U_{N\alpha a}\nu_a \longleftrightarrow \nu_a = \sum_{\alpha} U^*_{N\alpha a}\nu_\alpha
\label{am164}
\end{eqnarray}
The corresponding matrix for the charged leptons will be denoted by $U_E$.
The PMNS matrix is then given by the product
\begin{eqnarray}
V_{PMNS}=U_N^{\dagger}U_E
=\begin{bmatrix}
\langle \nu_e | \nu_1 \rangle & \langle \nu_e | \nu_2 \rangle & \langle \nu_e | \nu_3 \rangle\\
\langle \nu_\mu | \nu_1 \rangle & \langle \nu_\mu | \nu_2 \rangle & \langle \nu_\mu | \nu_3 \rangle\\
\langle \nu_\tau | \nu_1 \rangle & \langle \nu_\tau | \nu_2 \rangle & \langle \nu_\tau | \nu_3 \rangle
\end{bmatrix}
\begin{bmatrix}
\langle e_1 | e \rangle & \langle e_1 | \mu \rangle & \langle e_1 | \tau \rangle\\
\langle e_2 | e \rangle & \langle e_2 | \mu \rangle & \langle e_2 | \tau \rangle\\
\langle e_3 | e \rangle & \langle e_3 | \mu \rangle & \langle e_3 | \tau \rangle
\end{bmatrix}
\label{mm23cc}
\end{eqnarray}
where a mnemonic notation $\langle | \rangle$ has been used for the matrix elements of $U_N$ and $U_E$.
In the case at hand the mass eigenstates can be directly identified with the eigenvectors of the given eigenvalue problem tables \ref{tab331} and \ref{tab3322}. For the special values given after eq. (\ref{mm15}) one obtains
\begin{eqnarray}
\nu_e&=&(-0.034 - 0.49i, 0.49, 0.072 - 0.076i, -0.034 - 0.49i, 0.49, 0.072 - 0.076i) \nonumber \\
e&=&(0,0,-0.707,0,0,0.707) \nonumber \\
\nu_\mu&=&(0.078 + 0.47i, 0.476, -0.166 - 0.14i, 0.078 + 0.47i, 0.476, -0.166 - 0.14i) \nonumber \\
\mu &=&(0.500i, 0.500, 0, -0.500i, -0.500, 0) \nonumber \\
\nu_\tau &=& (0.066 + 0.158i, 0.066 - 0.158i, 0.665, 0.066 + 0.158i, 0.066 - 0.158i, 0.665)\nonumber \\
\tau&=&(0.500i, -0.500, 0, -0.500i, 0.500, 0)
\label{mm23}
\end{eqnarray}
where the 6-dimensional notation of eqs. (\ref{mm23h}) has been used.
According to eq. (\ref{njl4baa}) the isospin vector $\vec S$ gives the left-handed particle contribution, while $\vec T$ does the same for the righthanded anti-particles. Therefore the transformation from the neutrino and charged lepton mass eigenstates to their interaction eigenstates is given by the projection of the eigenvectors (\ref{mm23}) to $\vec S$ and $\vec T$, respectively.
Carrying out these operations one obtains
\begin{eqnarray}
V_{PMNS}=
\begin{pmatrix}
-0.627 + 0.144 i & 0.158 + 0.542 i & 0.141 + 0.500 i \\
0.012 + 0.311 i & 0.410 + 0.158 i & -0.831 - 0.141 i \\
0.047 + 0.697 i & -0.698+0.001i & -0.101 + 0.108 i
\end{pmatrix}
\label{ccjj1}
\end{eqnarray}
There is a CP violating effect in this matrix, because the Jarlskog invariant\cite{jarl} $J_{CP}$ is non-zero and given by $J_{CP}=0.0222$.
The point to note here is that large mixing angles for the PMNS matrix appear quite naturally, because the neutrino states lie near the total angular momentum $\vec \Sigma=\vec S+\vec T$ which is quite far away from the projection operator $\vec S$. In contrast, due to the dominance of the inter-tetrahedral interactions eq. (\ref{mm3dm}) the top quark state lies very close to the projection vector $\vec S$. In addition, this dominance will align all other quark flavors, thus forcing their CKM elements to small values.
On the quantitative side the result eq. (\ref{ccjj1}) unfortunately does not really reproduce the observed mixing angles for the neutrinos\cite{nuemix,altamix}.
A complete scan of the full parameter space seems unavoidable and should eventually better be carried out for the complete 24$\times$24 eigenvalue problem. This effort will be undertaken in future work.
\section{Conclusions}
In the preceeding sections a microscopic model for the SM Higgs mechanism has been applied to determine the quark and lepton masses and mixing angles. A discrete tetrahedral structure within a dynamical internal space has managed to fill the gap between the phenomenological hierarchy of mass scales. The underlying physical picture is that the universe resembles a huge crystal of internal molecules, each 'molecule' of tetrahedral form and arranged in such a way that certain symmetries are (spontaneously) broken. For such a model to be consistent, a (6+1)-dimensional space time has been introduced in ref.\cite{laa}, i.e. the 'molecules' extend to 3 internal dimensions orthogonal to physical space, and they interact via a (6+1)-dimensional QED featuring the necessary 'iso-magnetic' forces. The present-day structure of the universe is that of a (3+1)-dimensional 'surface crystal' built from an infinite set of tetrahedrons and living within the (6+1)-dimensional space time. For reasons of symmetry, no growth of the crystal is allowed into the internal dimensions.
The strong correlations within this system provide for the Higgs particle and the weak vector bosons as bound states.
Furthermore, internal spin excitations turn out to generate the correct quark and lepton spectrum. Then, it happens that an excitation in one internal tetrahedron is able to excite an excitation in the neighboring internal space and thus can travel as a quasi-particle through Minkowski space with a certain wave vector $\vec k$ which is to be interpreted as the physical momentum of the quark or lepton.
In this model, the SM symmetry breaking can be understood to proceed in 2 steps:
\begin{itemize}
\item the formation of a tetrahedron due to a new internal interaction force, which is 'antiferromagnetic' at small distances and leads to a frustrated configuration of isospin vectors fig. 1.
The frustrated tetrahedron breaks the internal spin-SU(2) as well as internal parity and time reversal to the Shubnikov group $A_4+S (S_4 - A_4)$. This symmetry breaking is not spontaneous but arises from the internal antiferromagnetic exchange interaction which avoids parallel spin states. The resulting local ground state fig. 1 is a chiral configuration, i.e. it violates internal and (as shown in ref.\cite{laa}) external parity, and the whole system is left $SU(2)_L$-symmetric in the following sense:
\item each local tetrahedral ground state can rotate independently of the others, i.e. it can freely rotate as a rigid body over its base point in Minkowski space, and this rotational symmetry of the rigid chiral spin vector system corresponds to a $SO(3)$ symmetry group, whose covering group is taken to define $SU(2)_L$. As a matter of fact it is a local symmetry, because the rotation can be different for tetrahedrons over different base points. Furthermore, due to the $V-A$ structure of the interactions induced by the chiral tetrahedral structure, it is a symmetry involving only left handed particles\cite{laa}. At large distances of the order of the Fermi scale the new interactions are (internally) ferromagnetic in nature and give rise to the global ferromagnetic order shown in fig. 2. Finally, the non-vanishing vev for the Higgs field $\sim \bar \psi \psi$ is due to a pairing mechanism as described in ref.\cite{laa}.
\end{itemize}
In order to analyze the mass problem of the fermions within this model, the most general $SU(2)_L\times SU(2)_R$ invariant NJL Lagrangian has been used\cite{njl2} as an effective approximation. Afterwards, a Heisenberg Hamiltonian for the internal spin vector interactions has been derived from that Lagrangian. This is justified because at the stage when the internal tetrahedron is formed chiral symmetry is still valid, so that one can describe the internal spin vibrations in terms of the chiral spin vectors $\vec S$ and $\vec T$.
Concerning the 'ferromagnetic' attraction between different tetrahedrons at large distances responsible for the SSB, it was noticed in section 5 that the gauge structure enforces an additional term which resembles the so called Dzyaloshinskii-Moriya interactions\cite{dmgut} of solid state physics. This term was interpreted quite naturally as a curvature effect of the $SU(2)_L$ gauge fields induced in the fiber bundle formed by the system of all tetrahedrons. The DM term simply takes care of this curvature to effectively maintain the 'ferromagnetic' order fig. 2.
Based on the prescribed model expressions for the quark and lepton mass spectrum were derived. It turned out that the extreme hierarchy in this spectrum can be attributed to the fact that the masses of different fermions get contributions from physically different sources, namely
\begin{itemize}
\item the top mass is dominated by a contribution of order $\Lambda_F$ which stems from the SSB {\it inter}-tetrahedral DM interactions. Physically it arises because the top quark corresponds to the 3 eigenmodes which 'disturb' the global ground state in the strongest possible way. This disturbance is also responsible for the hierarchy observed in the CKM matrix elements.
\item strange-, charm- and muon-mass are dominated by antiferromagnetic exchange couplings within one tetrahedron, and thus can be obtained from the {\it inner}-tetrahedral Heisenberg exchange couplings alone.
\item down-quark, up-quark and electron get their relatively small masses from energetically favored torsion contributions, which only concern 'radial' excitations of the internal spin vectors.
\item neutrino masses are protected by symmetry, because they correspond to vibrations in the valleys of the potential where all Heisenberg and even most of the torsional energy contributions vanish. This was shown to be related to a Goldstone effect associated to the breaking of the internal rotational SO(3) to the tetrahedral symmetry. While in ordinary ferromagnets after magnetization a U(1) symmetry about the z-axis survives, in the given frustrated configuration fig. 1 all three SO(3) generators give rise to Goldstone bosons, to be identified as the internal magnons corresponding to the 3 neutrino species.
\end{itemize}
Furthermore, the question of quark and lepton mixing was considered, albeit not in a very elaborate way.
The quark mixing is a complicated 24$\times$24 eigenvector problem with many parameters and a detailed analysis therefore postponed to future work. An attempt to determine the lepton mixing parameters was made. It turned out that the phenomenological values for the PMNS mixing angles cannot be obtained in a straightforward manner. The upshot of the discussion presented in this paper is that an accommodation to the measured neutrino properties is a non-trivial calculational exercise, because a complete scan over the parameter space of the anisotropic torsional couplings is needed.
To summarize, the quark and lepton masses have been successfully reduced to couplings among the internal spin vectors. Using these results the poor man's strategy (applied in this paper) is to choose the couplings so that the fermion masses and mixings come out right. The reader may rightfully object that everything done here is to replace one set of free parameters by another set, and one effective theory (the Standard Model) by another one (the NJL inspired internal Heisenberg model). However, as shown in ref.\cite{laa} the internal couplings can be calculated from first principles as exchange integrals over internal space just as in ordinary magnetism the exchange couplings of the Heisenberg model are in principle calculable from exchange integrals of electronic wave functions over physical space. A more ambitious program therefore is to determine all internal couplings from a fundamental theory like the higher dimensional QED considered in \cite{laa}.
|
\chapter{Knot Candidate Images}
\label{appendix:knotcandidateimages}
This appendix contains images of every knot candidate, in all optical bands for which we have data (F160W, F814W, F475W). These images differ from the rest of the images in this work, as they have a \emph{linear} stretch function applied, rather than a \emph{logarithmic} stretch function (the much smaller field-of-view requires a much smaller dynamic range). The scales of these images do not necessarily correspond; these images are primarily to demonstrate the shape and the level of agreement between the data contours, in red, and the contours of the fitted 2d-gaussian models, in green (the contours for both are fixed to identical levels). For all of the images, the jet runs from left to right.
\begin{figure}[p]
\centering
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f160w_32arcsec-eps-converted-to.pdf}
\caption{F160W}
\label{fig:knot_32arcsec_f160w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f814w_32arcsec-eps-converted-to.pdf}
\caption{F814W -- Possible point source}
\label{fig:knot_32arcsec_f814w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f475w_32arcsec-eps-converted-to.pdf}
\caption{F475W -- Non-detection}
\label{fig:knot_32arcsec_f475w}
\end{subfigure}
\caption{Knot located at 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}}
\label{fig:knot_32arcsec}
\end{figure}
\begin{figure}[p]
\centering
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f160w_43arcsec-eps-converted-to.pdf}
\caption{F160W}
\label{fig:knot_43arcsec_f160w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f814w_43arcsec-eps-converted-to.pdf}
\caption{F814W}
\label{fig:knot_43arcsec_f814w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f475w_43arcsec-eps-converted-to.pdf}
\caption{F475W}
\label{fig:knot_43arcsec_f475w}
\end{subfigure}
\caption{Knot located at 43\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}}
\label{fig:knot_43arcsec}
\end{figure}
\begin{figure}[p]
\centering
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f160w_106arcsec-eps-converted-to.pdf}
\caption{F160W}
\label{fig:knot_106arcsec_f160w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f814w_106arcsec-eps-converted-to.pdf}
\caption{F814W}
\label{fig:knot_106arcsec_f814w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f475w_106arcsec-eps-converted-to.pdf}
\caption{F475W}
\label{fig:knot_106arcsec_f475w}
\end{subfigure}
\caption{Knot located at 106\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}}
\label{fig:knot_106arcsec}
\end{figure}
\begin{figure}[p]
\centering
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f160w_112arcsec-eps-converted-to.pdf}
\caption{F160W -- Diagonal cutoff is from the edge of the detector}
\label{fig:knot_112arcsec_f160w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f814w_112arcsec-eps-converted-to.pdf}
\caption{F814W}
\label{fig:knot_112arcsec_f814w}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f475w_112arcsec-eps-converted-to.pdf}
\caption{F475W}
\label{fig:knot_112arcsec_f475w}
\end{subfigure}
\caption{Knot located at 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}}
\label{fig:knot_112arcsec}
\end{figure}
\clearpage
\newpage
\chapter{Future Work}
\label{appendix:future}
The study of Pictor A continues to confound the typical models used to explain AGN and jet emission. This suggests there is something to be learned, something beyond the most popular models of the physical processes within jets. While further analyses will continue in preparation for a journal publication of these results, there are key questions which go beyond the scope of this work. In order to better understand the perplexing Pictor A, we suggest future analysis which could be done on existing data (Section \ref{future:tidaltail}) and new observations which could add significant data (Section \ref{future:observations}).
\section{Tidal Tail Analysis}
\label{future:tidaltail}
These observations discovered a previously unknown and unexpected tidal tail (this tidal tail is the broad, sweeping feature North of Pictor A, most clearly visible in the F160W image; see Figure \ref{fig:galaxysubtractions_largefont}). This feature suggests a previous merger event in the history of Pictor A, and it might provide a history of the AGN and jet which we are studying. One hypothesis is that galaxy merger could play a role in triggering active galaxies; while this hypothesis does not appear to explain most AGN \cite{Villforth2014,Cisternas2011}, Pictor A could provide information about those AGN for which a merger might have been significant.
\section{Suggested Observations}
\label{future:observations}
As noted in Section \ref{significance:summary}, our ability to constrain emission models of Pictor A's jet remains largely limited by the available observational data. We suggest future observations which could break some of the degeneracies left by the current data.
\subsection{New Spectral Bands}
\label{future:observations:newspectralbands}
While observations stretch from radio to optical to x-ray bands, we are still missing crucial information for constraining synchrotron parameters. In particular, it would be useful to know the spectral index at radio frequencies, since such a determination could provide significant evidence for 2 distinct electron populations, if radio and x-ray spectral indices disagree.
Our synchrotron models predict radio emission to become brighter with increasing frequency, leading us to suggest the Atacama Large Millimeter/sub-millimeter Array (\emph{ALMA}). ALMA has recently begun operating with spectral bands as high as 900 GHz, compared to our current radio map taken at about 5 GHz. Furthermore ALMA is well-positioned at a latitude of 23\arcdeg{} South for observing Pictor A, which lies at a declination of $45\arcdeg$ South (outside the range of most Northern Hemisphere observatories). ALMA could leverage the brighter emission at higher frequencies to break key degeneracies in our spectral models of Pictor A's jet.
\subsection{Deeper Observations}
\label{future:observations:deeperobservations}
Deeper HST observations, while not breaking the degeneracy of radio spectral indices, might generate a greater number of knot detections. Many of the observed knot candidates (particularly those at 43\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} and 106\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}) were only marginally detected. Furthermore, the knot candidate at 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} appears to have a high flux (even in the existing F475W image), but in the current images, this knot lies partially beyond the edge of the F160W detector. Using the information we have gained, the pointing of the HST could easily be adjusted to capture that knot fully, on every detector. By fitting what we can see of the knot in the F160W image, it appears that it has a higher synchrotron cutoff than the other observed knots (compare the 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} SED in Figure \ref{fig:SED_nomodels_112arcsec} with the SED for the other observed knots in Figure \ref{fig:SED_nomodels_allknots}; see the images of the 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot in Figure \ref{fig:knot_112arcsec}). Not only does this mean the knot 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the core could yield different information about the jet, but it also means we are very likely to make a strong detection in all bands, given the high flux being produced.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=.8\linewidth]{SED_112arcsec-eps-converted-to.pdf}
\end{center}
\caption{SED of data only from the 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot. The F160W data in particular should be treated cautiously -- by trying to fit a 2d-gaussian, with only half of the profile visible (see Figure \ref{fig:knot_112arcsec_f160w}) we introduce potentially large, uncharacterized systematic errors. }
\label{fig:SED_nomodels_112arcsec}
\end{figure}
\chapter{Theory}
\label{theory}
In this chapter, I will summarize (and when necessary, derive) the key theoretical frameworks guiding this analysis. Chapter \ref{methodology} will discuss the implementation of the theory discussed in this chapter.
\section{Galaxy Subtraction}
In order to understand what was unusual about this galaxy, we first needed to be able to discriminate between what was normal and what was abnormal. In order to do that, we fit standard galaxy morphologies to our data, and analyzed the residual. Key to this process was understanding common galaxy morphologies, and how those morphologies might relate to our particular galaxy.
\subsection{Galaxy Morphologies}
\label{theory:morphology}
In the 1920s, Hubble noticed that galaxies can often be grouped into two main categories: ellipticals and disks, in a classification now known as the \emph{Hubble Sequence} \cite{HubbleSequence}. The specific surface brightness (as a function of radius), $\Sigma_\nu(r)$, of these respective categories are commonly fit by two profiles:
\begin{equation}
\Sigma_\nu(r) \sim
\begin{cases}
\exp\left(-\left[\frac{r}{r_0}\right] \right) & \mathrm{disk} \\
\exp\left(-\left[\frac{r}{r_0}\right]^{1/4} \right) & \mathrm{elliptical} \\
\end{cases}
\label{eq:SurfaceBrightnessProfiles}
\end{equation}
In 1963 S\'{e}rsic generalized these profiles, and developed the \emph{S\'{e}rsic index} \cite{SersicIndex}, which now is used to quantitatively describe galactic morphology. The surface brightness profile for an arbitrary S\'{e}rsic index, $n$, is:
\begin{equation}
\Sigma_\nu(r) \sim \exp\left(-\left[\frac{r}{r_0}\right]^{1/n} \right)
\label{eq:SersicProfile}
\end{equation}
The versatility of the S\'{e}rsic profile results in its usefulness as a generic galaxy morphology model. In particular, when looking at atypical aspects of a galaxy, a S\'{e}rsic profile can be subtracted to leave only the unusual aspects as residuals \cite{GALFIT}. Multiple S\'{e}rsic profiles (with different S\'{e}rsic indices, magnitudes, and scale radii, $r_0$) can then be combined to provide a useful, if incomplete basis set of models for describing a galaxy.
\section{Radiative Transfer}
\label{theory:radiativetransfer}
The following sections will summarize the field of high energy radiative transfer of non-thermal emission, as it applies to relativistic jets. In some cases, a more thorough handling is more appropriately left to review articles and books; in those cases, I refer the reader to appropriate sources. What is presented is a summary of key points, and derivations for aspects not typically discussed. In general, I will try to present equations in dimensionless form (following the style of \cite{Gould1979}), but when that is not possible, it is safe to assume Gaussian cgs units (e.g. spectral flux density, $S_\nu$, has units of erg s$^{-1}$ cm$^{-2}$ Hz$^{-1}$)
Before beginning, we assume a homogeneous plasma of charged particles of energies $E = \gamma m c^2$ (for a distribution of $\gamma$), moving through a constant magnetic field, $B$, with a pitch angle, $\alpha_{pitch}$, between their momenta and the local magnetic field. (We will refer to energies and Lorentz factors, $\gamma$, interchangeably, as they are trivially related by $E = \gamma m c^2$; it is easier to think of energy, but easier to express $\gamma$, which is independent of your choice of units).
We can assume electrons (and positrons) are responsible for a majority of the observed radiation, given their low mass (allowing for high accelerations, required to produce non-thermal radiation). For a number density of electrons, $n_e$, their distribution of energies is typically chosen to be a power law distribution with electron energy index $p$, and with low and high energy cutoffs imposed:
\begin{equation}
n_\gamma \equiv \frac{d n_e}{d \gamma} =
\begin{cases}
\kappa_e \gamma^{-p} & \gamma_\mathrm{min} < \gamma < \gamma_\mathrm{max} \\
0 & \mathrm{else}
\end{cases}
\label{eq:ElectronDistributionWithCutoffs}
\end{equation}
We may now begin to look at the predictions made by the hypothesized emission mechanisms: synchrotron radiation and inverse Compton up-scattering of background photons.
\subsection{Synchrotron Radiation}
\label{theory:synchrotron}
A free charge moving non-relativistically through a magnetic field will experience a Lorentz force:
\begin{equation}
\mathbf{F} = q(\mathbf{E} + \frac{\mathbf{v}}{c} \times \mathbf{B})
\label{eq:LorentzForce}
\end{equation}
In a constant magnetic field, the general solution for a freely moving charge governed by Equation \ref{eq:LorentzForce} is a helix, with a pitch angle, $\alpha_{pitch}$, between the magnetic field and the particle's velocity. It can then be shown that circular motion is described by $\alpha_{pitch} = \frac{\pi}{2}$ and linear motion is described by $\alpha_{pitch} = 0$. Under these conditions, a charge $q$ of mass $m$ will have a gyro-frequency (and thus an emission frequency) of:
\begin{equation}
\nu_g = \frac{q B}{2 \pi m c}
\label{eq:gryofrequency}
\end{equation}
When treated relativistically (with $\gamma = \frac{E}{m c^2}$), we find the more general fundamental frequency:
\begin{equation}
\nu_r = \gamma \nu_g \sin\left( \alpha_{pitch}\right)
\label{eq:synchrotronfrequency}
\end{equation}
with higher order harmonics (derived in \cite{LongairBook2ed}) leading to a maximal emission peak for a relativistic electron at a frequency:
\begin{equation}
\nu_{max} = .42 \gamma^2 \nu_g \sin\left( \alpha_{pitch}\right)
\label{eq:synchrotronfrequencymax}
\end{equation}
In astrophysical contexts though, we are often interested in an ensemble of electrons of different energies and pitch angles. The key results are well known (see \cite{LongairBook2ed,RybickiLightmanBook,BlumenthalGould1970}), but best summarized in \cite{Gould1979} under the assumptions of an isotropic spherically symmetric ensemble of electrons obeying the energy distribution in Equation \ref{eq:ElectronDistributionWithCutoffs}. (While Gould allows for a inhomogeneous source in \cite{Gould1979}, we will restrict ourselves to a homogeneous source for simplicity). Under those conditions, we expect an (optically thin) flux density spectrum:
\begin{equation}
S_\nu = \frac{2 \pi \hbar}{3}\nu \rho_\nu \frac{R^3}{d_L^2}
\label{eq:gouldfluxspectrum}
\end{equation}
for a source of radius, $R$, at a luminosity distance, $d_L$, and which has a specific volumetric production rate of photons:
\begin{equation}
\rho_\nu = 4 \pi \left(\frac{3}{2}\right)^\frac{p-1}{2} a\!\left(p\right) \alpha_{FS} \kappa_e \left( \frac{\nu_g}{\nu}\right)^\frac{p+1}{2}
\label{eq:gouldrhonu}
\end{equation}
where $\alpha_{FS} \equiv e^2 / h c$ is the fine structure constant, and $a(p)$ is given in \cite{Ginzburg65} to be:
\begin{equation}
a(p) = \frac{2^\frac{p -1}{2}}{8 \left(p+1\right)} \sqrt{\frac{3}{\pi}} \frac{\Gamma\left( \frac{3p -1}{12} \right) \Gamma\left( \frac{3p +19}{12} \right) \Gamma\left( \frac{p +5}{4} \right)}{\Gamma\left( \frac{p+7}{4} \right)}
\label{eq:BlumenthalGoulda}
\end{equation}
leading to a power law spectrum:
\begin{equation}
S_\nu \sim \nu^{-\frac{p-1}{2}} = \nu^{- \alpha_{sync}}
\label{eq:synchrotronunabsorbedpowerlaw}
\end{equation}
which defines the synchrotron spectral index:
\begin{equation}
\alpha_{sync} \equiv \frac{p-1}{2}
\label{eq:synchrotronindex}
\end{equation}
It is important to note the range of validity of Equation \ref{eq:synchrotronunabsorbedpowerlaw}. The derivation assumed an unbounded electron power law (which is non-normalizable) -- that approximation is good as long as we are far from significant contributions from the bounds of our electron power law: $\gamma_{min}^2 \nu_g \ll \nu \ll \gamma_{max}^2 \nu_g$. At high frequencies, Equation \ref{eq:synchrotronfrequencymax} predicts a cutoff at:
\begin{equation}
\nu_{max} \approx \gamma_{max}^2 \nu_g
\label{eq:synchrotronnumaxgammamax}
\end{equation}
A similar cutoff exists at low frequencies, but more commonly we find that the low frequency departure from Equation \ref{eq:synchrotronunabsorbedpowerlaw} is due to the source becoming optically thick, a possibility for which we did not account.
\subsubsection{Synchrotron Self-Absorption}
\label{theory:synchrotronselfabsorption}
Properly accounting for optical depth (see \cite{LongairBook3ed} for a detailed derivation) we expect a specific optical depth, $\tau_\nu$, after passing through a length, $\ell$, of the form \cite{Gould1979}:
\begin{equation}
\tau_\nu = c(p)r_e^2 \kappa_e \frac{\nu_e}{\nu} \left( \frac{\nu_g}{\nu} \right)^\frac{p+2}{2} \ell
\label{eq:synchrotronopticaldepth}
\end{equation}
where $r_e = e^2 / m_e c^2$ is the standard classical electron radius $\nu_e = c / r_e$ is a less standard definition used to simplify the equation and its units, and $c(p)$ is given by \cite{Gould1979} to be:
\begin{equation}
c = \frac{3^{\frac{p+1}{2}}}{8} \sqrt{\pi} \frac{ \Gamma\left(\frac{p+6}{4}\right)\Gamma\left(\frac{3p+2}{12}\right)\Gamma\left(\frac{3p+22}{12}\right)}{\Gamma\left(\frac{p+8}{4}\right)}
\label{eq:Gouldc}
\end{equation}
The process of a source both emitting, then reabsorbing synchrotron radiation is called \emph{synchrotron self-absorption} (\emph{SSA}). Equation \ref{eq:synchrotronopticaldepth} predicts that this is more likely to dominate at low frequencies. This defines a frequency, $\nu_{SSA}$ (such that $\tau_{\nu_{SSA}} = 1$), determining the transition between the low-frequency optically thick regime ($\tau_\nu > 1$) and the high-frequency optically thin regime (following Equation \ref{eq:synchrotronunabsorbedpowerlaw}). This transition frequency can be found using Equation \ref{eq:synchrotronopticaldepth}:
\begin{equation}
\nu_{SSA} = \left( c(p) r_e^2 \kappa_e \nu_e \nu_g^\frac{p+2}{2} R \right)^\frac{2}{p+4}
\label{eq:nuSSA}
\end{equation}
For a more complete discussion of the effect of a specific optical depth, $\tau_\nu$, see \cite{LongairBook3ed}, but ultimately it modifies the spectrum given in Equation \ref{eq:synchrotronunabsorbedpowerlaw} to one that has the limiting behaviors:
\begin{equation}
S_\nu \sim
\begin{cases}
\nu^{5/2} & \nu_g \gamma_{min}^2 \ll \nu \ll \nu_\mathrm{SSA} \\
\nu^{-\frac{p + 1}{2}} & \nu_{SSA} \phantom{\gamma} \ll \nu \ll \nu_g \gamma_{max}^2 \\
\end{cases}
\label{eq:SynchrotronSpectrumFull}
\end{equation}
This spectral shape, can be seen in Figure \ref{fig:synchrotronspectrum}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=.7\linewidth]{SynchrotronSpectrum-eps-converted-to.pdf}
\end{center}
\caption{Sample synchrotron spectrum, with asymptotic behaviors marked. For optically thick frequencies ($\nu \ll \nu_{\mathrm{SSA}}$) we expect $S_\nu \sim \nu^{5/2}$, and for the optically thin frequencies ($ \nu \gg \nu_\mathrm{SSA}$) we expect $S_\nu \sim \nu^{-\frac{p-1}{2}}$.}
\label{fig:synchrotronspectrum}
\end{figure}
\subsubsection{Energetics}
\label{theory:synchrotronenergetics}
In synchrotron emission, energetics provide a convenient way to connect astronomical observables and physical conditions at the source.
The maximal frequency cutoff, $\nu_{max}$ (given in Equation \ref{eq:synchrotronnumaxgammamax}) can set limits on the lifetime of the source. Not only does Equation \ref{eq:synchrotronfrequencymax} connect the observed frequency cutoff to an inferred electron energy cutoff, but it also predicts that higher energy electrons will radiate energy at a greater rate, depleting their energy more rapidly. Solving for the time evolution of this system (see \cite{Ginzburg65}) predicts that after a time, $t$, there should be effectively no electrons above a certain energy threshold:
\begin{equation}
\gamma_{max}(t) < \frac{9}{8 \pi} \frac{1}{\alpha_{FS}} \frac{m_e c^2}{h \nu_g} \frac{1}{\nu_g t}
\label{eq:electronenergylimit}
\end{equation}
which equivalently places a limit on the time since the most recent electron re-acceleration event:
\begin{equation}
t < \frac{9}{8 \pi} \frac{1}{\alpha_{FS}} \frac{m_e c^2}{\gamma_{max} h \nu_g} \frac{1}{\nu_g}
\label{eq:electronagelimit}
\end{equation}
If we assume during that time, $t$, the system has sought a state of minimum total energy (summing kinetic energy of particles in the jet and energy stored in the magnetic field; see Figure \ref{fig:minimumenergy}) we can also determine the minimum-energy magnetic field, $B_{me}$, for a given observation \cite{Worrall2006}:
\begin{figure}[tb]
\begin{center}
\includegraphics[width=.7\linewidth]{Bme_longair-eps-converted-to.pdf}
\end{center}
\caption{Sample curve demonstrating the total energy of a synchrotron emitting system, and the magnetic field, $B_{me}$ (marked $B_{min}$ above; see Equation \ref{eq:BmeWorrall} for analytic form), satisfying the minimum energy criterion. Figure taken from \cite{LongairBook3ed}.}
\label{fig:minimumenergy}
\end{figure}
\begin{equation}
B_{me} = \left[ \frac{ (p+1) C_1}{4 C_2} \frac{1 + K}{\eta} \frac{4 \pi d_L^2 (1+z)^\frac{p-3}{2}}{V} S_\nu \nu^\frac{p-1}{2} \frac{ \left( \gamma_{max}^{2-p} - \gamma_{min}^{2-p} \right) }{2-p} \right]^\frac{2}{(p + 5)}
\label{eq:BmeWorrall}
\end{equation}
where $K$ is the ratio of kinetic energy of electrons to the kinetic energy of other particles ($K \approx 0$ gives a ``true'' minimum configuration), $\eta$ is the filling fraction of the emitting volume relative to the relevant volume containing magnetic fields ($\eta =1$ corresponds to no voids, also used for finding a ``true'' minimum), $V$ is simply the volume of the emitting region, and the constants $C_1$, $C_2$ are implicitly defined as:
\begin{eqnarray}
C_1 &\equiv& \left( \frac{\nu_g}{B} \right)^{-\frac{p-1}{2}} \frac{c}{r_e} \frac{p+1}{3^{p/2} \pi^{3/2}} \frac{\Gamma\left( \frac{p +7}{4}\right)}{\Gamma\left( \frac{p +5}{4}\right)\Gamma\left( \frac{3p +19}{12}\right)\Gamma\left( \frac{3p -1}{12}\right)}
\label{eq:WorrallC1} \\
C_2 &\equiv& \frac{u_{mag}}{B^2} = \frac{1}{8 \pi} \:\mathrm{(for \: Gaussian \: cgs)}
\label{eq:WorrallC2}
\end{eqnarray}
The validity of the assumption of a minimum energy configuration can best be tested through x-rays observations, and an understanding of the information inferred about inverse Compton radiative processes \cite{Worrall09}. In general, such observations and understandings often, but not always, provide good evidence for a roughly minimal energy configuration \cite{Worrall2006}. To further characterize our system, we must now move on to the theory of \emph{inverse Compton scattering} (\emph{IC}).
\subsection{Inverse Compton Scattering}
\label{theory:InverseComptonScattering}
Another significant source of radiation coming from relativistic plasmas is inverse Compton scattering of photons. This inelastic scattering is independent of the local magnetic field (in contrast to synchrotron radiation) but requires a source of \emph{seed} photons. In particular, we'll consider two sources of these photons: external photons (predominantly the cosmic microwave background, \emph{CMB}) and internal photons (such as the source's own synchrotron emission). We will focus mainly on the analytic theory of external inverse Compton scattering, which will provide a framework that is used numerically to compute synchrotron self-Compton scattering (\emph{SSC}).
Compton scattering is an inelastic scattering process which allows for transfer of momentum and energy between photons and electrons. In the electron's rest frame, it appears as if an incoming photon has given momentum to the previously stationary electron, increasing the electron's energy. This process appears qualitatively different when viewed in the center of momentum frame for an ensemble of approximately isotropic electrons and photons. If the electrons have enough energy, it will appears as if the electrons give energy to the photons. In \emph{Compton scattering} (in the electron's rest-frame) it appears that the photon loses energy, but in \emph{inverse Compton scattering} (the relativistically equivalent process viewed from the center-of-momentum frame) it appears that the photon has gained energy.
The key technical points are how much energy gets transferred, and often the scattering takes place. The first is relatively easy to answer. A low energy photon of frequency $\nu_0$, scattering off an electron of energy $E = \gamma m c^2$, will, on average, result in a photon of frequency:
\begin{equation}
\nu \approx \frac{4}{3} \gamma^2 \nu_0
\label{eq:ComptonBoosting}
\end{equation}
The rate at which this happens depends on the density of photons, the density of electrons, and the cross-section of the interaction.
\subsubsection{Cross-Sections}
\label{theory:InverseComptonCrossSections}
For a single electron, it is most convenient to compute this cross-section in the electron's rest frame. In that frame, the general result is the \emph{Klein-Nishina cross-section} (derived in \cite{LongairBook2ed}):
\begin{equation}
\sigma_{\mathrm{K-N}} = \pi r_e^2 \frac{1}{x} \left( \left[1 - \frac{2 \left( x+1 \right)}{x^2} \right] \ln \left( 2x + 1 \right) + \frac{1}{2} + \frac{4}{x} - \frac{1}{2 \left( 2x +1 \right)^2} \right)
\label{eq:KleinNishinaCrossSectionLongair}
\end{equation}
using the standard definition of the classical electron radius, $r_e \equiv \frac{e^2}{m_e c^2}$, and defining a dimensionless photon energy, $x \equiv \frac{h \nu}{m_e c^2}$. This can be simplified further in terms of the standard \emph{Thomson cross-section}:
\begin{equation}
\sigma_T = \frac{8 \pi}{3} r_0^2
\label{eq:ThomsonCrossSection}
\end{equation}
which gives us the Equation \ref{eq:KleinNishinaCrossSectionLongair} (in the form given by \cite{RybickiLightmanBook}):
\begin{equation}
\sigma_{\mathrm{K-N}} = \frac{3 \sigma_T}{4} \left( \frac{1 + x}{x^3} \left[\frac{2x \left(1 +x\right)}{1 + 2x} - \ln \left(1 + 2x \right) \right] + \frac{\ln \left(1 + 2x\right)}{2x} - \frac{1 + 3x}{\left(1 + 2x\right)^2} \right)
\label{eq:KleinNishinaCrossSectionRybickiLightman}
\end{equation}
which has the asymptotic behavior:
\begin{equation}
\sigma_{\mathrm{K-N}} \approx
\begin{cases}
\sigma_T (1 - 2x) \approx \sigma_T & x \ll 1 \\
\frac{3 \sigma_T}{8} \frac{1}{x} \left[ \ln \left(2x\right) + \frac{1}{2} \right] & x \gg 1
\end{cases}
\label{eq:KleinNishinaCrossSectionLimits}
\end{equation}
(It is important to remember that in the definition of $x \equiv \frac{h \nu}{m_e c^2}$, we care about the photon frequency in the \emph{electron's} rest frame. While a photon might have a frequency $\nu_0$ in the observer's frame, an electron with energy $E = \gamma m c^2$ could boost that frequency to $\nu = \gamma \nu_0$ in the electron's rest frame.)
Ultimately, we will find that only the Thomson regime plays a significant role in this work (high energy interactions are suppressed due to the $\sigma_\mathrm{K-N} \sim x^{-1}$ dependence at high energies). For that reason we will focus on the details of the low-energy Thomson regime of inverse Compton scattering, where the cross-section is approximately independent of energy. We leave discussions of the high-energy Klein-Nishina regime to other sources (see \cite{Jones1968,BlumenthalGould1970}).
\subsubsection{General Method for Thomson-Limit Inverse Compton Scattering}
\label{theory:InverseComptonGeneralMethod}
While the technical details of inverse Compton scattering have been extensively covered in other texts (e.g. \cite{BlumenthalGould1970}), we will introduce a simplified approach which we will use to derive results not typically discussed. The method we introduce has the ability to quickly determine the spectral shape of the scattered radiation, while providing a more intuitive view than the classic texts.
Using the Thomson cross-section, we find the general formula for the production of high energy photons of frequency, $\nu$ (shown through Equations 2.44, 2.45, 2.46 and 2.61 of \cite{BlumenthalGould1970}):
\begin{equation}
\frac{dn_\nu}{dt} = \int n_\gamma\; \sigma_\mathrm{T} \; c \; n_{\nu_0}\!\!\left(\!\nu_0 = \frac{3 \nu}{4 \gamma^2}\right) d\gamma
\label{eq:generalICscatteringrate}
\end{equation}
for a distribution of electron energies, $n_\gamma$ (e.g. Equation \ref{eq:ElectronDistributionWithCutoffs}), and an distribution of photon frequencies, $n_{\nu_0}$ (the momenta of each distribution are assumed to be isotropic).
To connect to terms more convenient for radiative transfer, we then find the emission coefficient:
\begin{equation}
j_\nu = \frac{h \nu}{4 \pi} \frac{dn_\nu}{dt} = \frac{h c \sigma_T}{4 \pi} \nu \int n_\gamma \; n_{\nu_0}\!\!\left(\!\nu_0 = \frac{3 \nu}{4 \gamma^2}\right) d\gamma
\label{eq:emissioncoefficientgeneral}
\end{equation}
If we assume a homogeneous, optically thin source (see Section \ref{theory:InverseComptonMultipleScatterings} for a relaxation of this assumption), we then find the relation for the shape of the specific flux density:
\begin{equation}
S_\nu \sim \nu \int n_\gamma \; n_{\nu_0}\left(\nu_0=\frac{3 \nu}{4 \gamma^2}\right) d\gamma
\label{eq:generalICflux}
\end{equation}
Just as the spectral shape of synchrotron emission was a key observable discussed in Section \ref{theory:synchrotron}, it will be a useful observable when discussing inverse Compton emission.
We can now use Equation \ref{eq:generalICflux}, and assumptions of a homogeneous source, to derive some basic properties of inverse Compton mechanisms within jets.
\subsubsection{Inverse Compton Scattering of the CMB}
\label{theory:InverseComptonCMB}
There are two key regimes when understanding inverse Compton scattering of the cosmic microwave background (\emph{IC-CMB}) in the Thomson limit. We first will deal with the low frequency tail, which obeys a Rayleigh-Jeans power law, and then we will deal with the higher frequency spectrum, which must take the spectral peak of the CMB black body into account.
\paragraph{Inverse Compton Scattering of a Power Law Photon Distribution}
\label{theory:InverseComptonCMBPowerLaw}
We will first consider the lowest frequencies of the produced spectrum:
\begin{equation}
\nu \ll \gamma_\mathrm{min}^2 \nu_\mathrm{0, peak}
\end{equation}
where $\nu_{0, \mathrm{peak}}$ is the frequency at the peak of the seed photon distribution. For a blackbody of temperature, \emph{T}, this peak frequency is given by Wien's law:
\begin{equation}
\nu_\mathrm{0, peak} = 5.9 \cdot 10^{10} \; T \; \frac{\mathrm{Hz}}{\mathrm{K}}
\label{eq:WienPeak}
\end{equation}
For those frequencies, we expect every seed photon to come from the Rayleigh-Jeans tail of the blackbody distribution ($\nu_0 \ll \nu_{0, peak}$):
\begin{equation}
n_{\nu_0} \sim \frac{S_{\nu_0}}{\nu_0} \sim \frac{\nu_0^2}{\nu_0} = \nu_0
\label{eq:RayleighJeans}
\end{equation}
Using Equations \ref{eq:RayleighJeans} and \ref{eq:generalICflux} we can then find the expected shape of the produced spectrum at low frequencies:
\begin{eqnarray}
S_\nu &\sim& \nu \int_{\gamma_\mathrm{min}}^{\gamma_\mathrm{max}} n_\gamma(\gamma) \; n_{\nu_0}\!\!\left(\! 3\nu / 4\gamma^2\right) d\gamma \nonumber \\
&\sim& \nu \int_{\gamma_\mathrm{min}}^{\gamma_\mathrm{max}} \gamma^{-p} \; \frac{\nu}{\gamma^2} d\gamma \nonumber \\
&\sim& \nu^2 \int_{\gamma_\mathrm{min}}^{\gamma_\mathrm{max}} \gamma^{-p-2} d\gamma \\
&\sim& \nu^2
\label{eq:InverseComptonRayleighJeans}
\end{eqnarray}
(recognizing that $\int_{\gamma_\mathrm{min}}^{\gamma_\mathrm{max}} \gamma^{-p-2} d\gamma $ is constant across all frequencies, determined by the index and bounds of the electron energy power law). At low frequencies, the spectral shape of the Rayleigh-Jeans tail $\left(S_\nu \sim \nu^2 \right)$ is preserved through the inverse Compton scattering.
Above $\nu = \nu_\mathrm{0, peak} \gamma_\mathrm{min}^2$, the Rayleigh Jeans approximation (Equation \ref{eq:RayleighJeans}) no longer holds. We must account for the peak of the blackbody spectrum.
\paragraph{Inverse Compton Scattering of a Monochromatic Photon Distribution}
\label{theory:InverseComptonCMBMonochromaticseed}
For frequencies, $\nu \gg \nu_\mathrm{0, peak} \gamma_\mathrm{min}^2$, the peak of the blackbody begins to have an effect on the produced spectrum. Due to the peaked nature of the blackbody spectrum, we now approximate the seed spectrum as monochromatic -- a delta function at $\nu_0 = \nu_\mathrm{0, peak}$.
This derivation is common, and exhaustive discussions can be found in \cite{RybickiLightmanBook,BlumenthalGould1970} (although it is worth noting that the approach using Equation \ref{eq:generalICflux} to find an analog of Equation \ref{eq:InverseComptonRayleighJeans} could also be used to reproduce this well-known result). Here we will only quote the resulting shape (valid at frequencies $\nu_{0,peak} \gamma_{min}^2 \ll \nu \ll \nu_{0, peak} \gamma_{max}^2$):
\begin{equation}
S_\nu \sim \nu^{-\frac{p - 1}{2}}
\label{eq:InverseComptonMonochromatic}
\end{equation}
Combining the theoretical results of the Rayleigh-Jeans power law component (\ref{eq:InverseComptonRayleighJeans})) and monochromatic component of the IC-CMB spectrum (\ref{eq:InverseComptonMonochromatic}), we then find:
\begin{equation}
S_\nu \sim
\begin{cases}
\nu^2 & \phantom{a \nu_\mathrm{0, peak} \gamma_\mathrm{min}^2 \ll } \nu \ll \nu_\mathrm{0, peak} \gamma_\mathrm{min}^2 \\
\nu^{-\frac{p - 1}{2}} & \nu_\mathrm{0, peak} \gamma_\mathrm{min}^2 \ll \nu \ll \nu_\mathrm{0, peak} \gamma_\mathrm{max}^2
\end{cases}
\label{eq:InverseComptonSpectrum}
\end{equation}
It is important to compare this spectral shape to that of the synchrotron spectrum including self-absorption, Equation \ref{eq:SynchrotronSpectrumFull}. Both start with a sharp initial rise, the shape of which is independent of the underlying electron distribution. Then, at higher frequencies, there is a kink, followed by identical photon indices ($S_\nu \sim \nu^{-\alpha}$):
\begin{equation}
\alpha_\mathrm{IC-CMB} = \frac{p - 1}{2} = \alpha_\mathrm{sync}
\end{equation}
\subsection{Synchrotron Self-Compton}
\label{theory:InverseComptonSynchrotronSelfCompton}
While in Section \ref{theory:InverseComptonCMB} we discussed inverse Compton scattering using the CMB as a seed photon field, many of the same principles can also be applied to inverse Compton scattering of the synchrotron emission produced. As this process requires the same ensemble of electrons to both produce, and then inverse Compton scatter the synchrotron radiation, it is termed \emph{synchrotron self-Compton} radiation (SSC).
The general process is to scatter all local photon fluxes (e.g. starlight, AGN emission, CMB, synchrotron emission, etc); far from any significantly luminous sources, the CMB will often be the dominant source of photons at the relevant radio frequencies. It is also conceivable that an optically thick source might even inverse Compton scatter photons which have previously undergone inverse Compton scattering.
\subsection{Multiple Scatterings \& Comptonisation}
\label{theory:InverseComptonMultipleScatterings}
Similar to our initial discussion of synchrotron emission, our discussion of inverse Compton scattering has assumed optically-thin inverse Compton scatterings. For Compton scattering in the Thomson limit (see Section \ref{theory:InverseComptonCrossSections}), we expect a specific optical depth:
\begin{equation}
\tau_\nu = n_e \sigma_T R
\label{eq:thomsonopticaldepth}
\end{equation}
where $n_e$ is the number density of electrons:
\begin{equation}
n_e = \int_{\gamma_{min}}^{\gamma_{max}} n_\gamma d\gamma = \kappa_e \int_{\gamma_{min}}^{\gamma_{max}} \gamma^{-p} d\gamma = \frac{\kappa_e}{1-p} \left( \gamma_{max}^{1-p} - \gamma_{min}^{1-p} \right)
\end{equation}
Ultimately, we will focus little on multiple scatterings, as we find they play little role ($\tau_\nu < 1$). For low energy electrons ($\gamma-1 \ll 1$) the topic of multiple scatterings characterized by the Kompaneets equation (original derived in \cite{Kompaneets1957}; see \cite{LongairBook3ed,RybickiLightmanBook} for suitable summaries). For high energy electrons though (as in our case), we expect the \emph{Compton catastrophe} to drain electron energy extremely rapidly if:
\begin{equation}
\eta \equiv \frac{L_{IC}}{L_{sync}} = \frac{ \frac{4}{3} \gamma^2 \sigma_T c u_{rad}}{\frac{4}{3} \gamma^2 \sigma_T c u_{mag}} = \frac{u_{rad}}{u_{mag}} > 1
\label{eq:synchrotroncatastrophe}
\end{equation}
where $u_{rad}$ is the energy density of the photons, and $u_{mag}$ is the energy density of the magnetic field. For more details see \cite{LongairBook3ed}.
\subsection{Bulk Doppler Boosting}
\label{theory:boosting}
In jets, it is not unusual for the previously described synchrotron, SSC and IC-CMB models to be insufficient for modeling observed fluxes. In particular, the assumption of an isotropic distribution of electrons, whose center-of-momentum frame is the same as the observer's frame, should be suspect since we are dealing with a system undergoing a bulk transfer of mass and power. If we transform into a frame moving with velocity $v = \beta c$ relative to the observer's frame, the electrons then might be isotropic, undergoing the previously outlined emission processes. That emission would then be affected by a bulk Lorentz factor $\Gamma = \frac{1}{\sqrt{1 - \beta^2}}$. When we take into account the angle, $\theta$, between the jet and the line-of-sight (and correspondingly define $\mu \equiv \cos{\theta}$), we get a bulk \emph{Doppler factor}:
\begin{equation}
\delta = \frac{1}{\Gamma \left(1 - \mu \beta \right)}
\label{eq:dopplerfactor}
\end{equation}
This bulk Doppler factor is quite noticeable in how it can boost the observed IC-CMB flux relative to the observed synchrotron flux \cite{Worrall09}. While the Doppler factor, $\delta$, is an observable, the inversion to $\mu$ or $\Gamma$ is degenerate. We can only choose a fixed $\mu$ and find the required $\Gamma$ (or vice versa) required to explain the observed ratio of synchrotron and IC-CMB fluxes. In particular, it's possible to find the extremal values for $\Gamma$ and $\mu$ which lead to physical values for the other.
Once an assumption for $\Gamma$ or $\mu$ has been made, solving for the other becomes simple using the formalism of \cite{Marshall2005}. Defining a ratio between synchrotron and Compton emission:
\begin{equation}
R_1 \equiv \frac{S_{\nu_{x-ray}} \nu_{x-ray}^{\alpha}}{S_{\nu_{radio}} \nu_{radio}^{\alpha}}
\label{eq:MarshallR1}
\end{equation}
where the ``1'' subscript denotes the value that our measurements suggest without the inclusion of Doppler boosting ($\Gamma =1$, hence the notation). We can apply that notation to the magnetic field ($B_{me} \approx B_1 = B \delta$ for $B_{me}$ from Equation \ref{eq:BmeWorrall}), and then group observables into a dimensionless parameter:
\begin{equation}
K = B_1 \left(a R_1\right)^{\frac{1}{\alpha + 1}} \left( 1 + z \right)^{-\frac{\alpha + 3}{\alpha + 1}}b^{\frac{1 - \alpha }{1 + \alpha}}
\label{eq:MarshallK}
\end{equation}
with the constants $a$ and $b$, which are given in Gaussian cgs units as:
\begin{eqnarray}
a &=& 9.947 \cdot 10^{10} \quad \mathrm{G}^{-2}
\label{eq:Marshalla} \\
b &=& 3.808 \cdot 10^{4\phantom{1}} \quad \mathrm{G}
\label{eq:Marshallb}
\end{eqnarray}
Finally, this allows us to solve for $\Gamma$ or $\theta$ (via $\beta$ or $\mu$ respectively) using Equation 4 of \cite{Marshall2005}:
\begin{equation}
K = \frac{1 - \beta + \mu - \mu \beta}{\left( 1 - \mu \beta \right)^2}
\label{eq:MarshallKbetamu}
\end{equation}
\subsection{Extensions of Synchrotron Radiation and Inverse Compton Scattering}
\label{theory:Extensions}
While the theory outlined above is a standard approach for modeling such a system, a more complicated model might be required to describe the emission. We will chiefly discuss additional components to the electron energy distribution, as well deviations from the assumption of isotropic pitch angles, $\alpha_{pitch}$.
We first note that there might be two components to the electron energy power law. The simplest case is a \emph{kink} in the power law, motivated partially by the electron energy lifetime estimates of Section \ref{theory:synchrotronenergetics} which would preferentially deplete the electron spectrum at high energies. A more general model would be a two-component electron energy distribution with a distinct low-energy population and a distinct high-energy population (rather than 2 populations which form one continuous energy distribution). Such a model could explain x-ray emission without appealing to inverse Compton scattering.
It also is possible the electron energy distribution does not follow a power law form. In particular, \cite{Allen2008} found observational evidence for a ``curved'' electron energy distributions in supernova remnants, which was inspired by theoretical modeling of the shock plasma. While the theory of the physical processes in jets is not the same as for those in supernova remnants, it does suggest that a more complete handling of the jet physics could yield qualitatively different results, if only we understood the processes within jets well enough.
Finally, the overall assumption of an isotropic distribution of electron momenta was convenient, but could be relaxed. If electron pitch angles, $\alpha_{pitch}$, aren't correlated with energy, then the distribution of pitch angles only sets the normalization of the synchrotron spectrum, and the location of a synchrotron self-absorption break. If the pitch angles do correlate with energy (electrons with larger pitch angles tend to radiate faster, so we would expect fewer high energy electrons with large pitch angles) then the modeling becomes significantly more complex, going beyond the scope of this work.
Many of these extensions are best handled numerically, rather than analytically. Such numerical models are built upon the frameworks which we have developed in this chapter.
\chapter{Methodology}
\label{methodology}
\section{HST Observations}
3 images were proposed for, and obtained, under the HST Guest Observer program (\emph{GO} proposal ID 12261 \cite{HSTProposal}). These three images were captured with the Wide Field Camera 3 (\emph{WFC3}; for technical details, see the instrument handbook \cite{WFC3InstrumentHandbook}). In order to gain optical spectral information, three wideband filters were used: F160W, F814W, F475W (centered approximately on 1600nm, 814nm, 475nm respectively; see Appendix A of the instrument handbook \cite{WFC3InstrumentHandbook} for filter details). The F160W image was taken on the infrared detector of WFC3 (\emph{WFC3/IR}), while the F814W and F475W images were taken on the UV-visible detector (\emph{WFC3/UVIS}). For more details on the chosen observational parameters, see Table \ref{table:observations}.
\begin{table*}[tbp]\centering
\ra{1.3}
\begin{tabular}{lrrrrrr@{}}\toprule
& & {F160W} & \phantom{abc}& {F814W} &
\phantom{abc} & {F475W}\\ \midrule
\emph{Instrument} \\
& WFC3 Detector & IR & & UVIS & & UVIS \\
& Native Resolution, [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .13 & & .04 & & .04 \\
& Drizzled Resolution, [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .07 & & .02 & & .02 \\
\emph{Filter} \\
& Pivot wavelength, [nm] & 1536 & & 802 & & 477 \\
& Spectral width, [nm] & 268 & & 153 & & 134 \\
\emph{General} \\
& Exposure Time, [s] & 2708 & & 1200 & & 1299 \\
& Dataset & IBJX01010 & & IBJX01030 & & IBJX01020 \\
\bottomrule
\end{tabular}
\caption{\label{table:observations} Overview of image parameters for relevant HST images. For more technical details on detectors, filters and other instrumental aspects, see \cite{WFC3InstrumentHandbook}. Otherwise all parameters of the observation are similar.}
\end{table*}
\section{HST Pipeline}
\label{methodology:pipeline}
The Space Telescope Science Initiative (\emph{STScI}), on behalf of the National Aeronautics and Space Administration (\emph{NASA}), has developed a standard data calibration pipeline for data taken by the HST. While for many purposes the default pipeline is sufficient, many users find it more effective to tweak the parameters of this pipelining process. In this section, we will discuss a few key aspects, as they pertain to this work.
\subsection{Drizzling}
\label{methodology:pipeline:drizzling}
Our work relied heavily on the technique of \emph{drizzling} to harness the full observational power of the HST. Drizzling is a process that takes multiple dithered exposures (images which are slightly offset in their pointing) and combines them to create an image of higher resolution than the native detector image resolution \cite{DrizzleMethod}. Drizzling assumes the image was undersampled -- the resolution must be limited by the detector, not by optical diffraction limits. As we were hoping to resolve jet substructure, proper drizzling of our images was key to providing the most angular information as possible.
The specific drizzling implementation used in this work was DrizzlePac, which at the time of this writing, was the software supported by the STScI for use with HST \cite{DrizzlePacHandbook}.
\paragraph{Geometric Distortion Correction}
\label{methodology:pipeline:drizzling:geometricdistortioncorrection}
As drizzling takes a series of exposures, in detector coordinates, and returns a reconstructed image in sky coordinates, it is also a convenient time to correct for geometric distortions inherent to the detector. Due to instrumental optical effects, the rays reaching the detector are not parallel in all regions \cite{WFC3DataHandbook} (see Figure \ref{fig:geometricdistortions} for a sample distortion map). This results in a squeezing or stretching of apparent angular sizes. These distortions are well-characterized, and the drizzling process can invert that distortion, ensuring that all pixels represent equal solid angles of the sky. The photometry required to determine fluxes involves integrating over solid angles of the sky, thus is it critical that solid angles are properly represented in our images.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=.8\linewidth]{wfc3_geometric_distortions_ir.jpg}
\end{center}
\caption{WFC3/IR geometric distortion map, taken from the WFC3 Data Handbook \cite{WFC3DataHandbook}. Map is in detector pixel coordinates, with crosses representing distortion of angular sizes (scalar distortions) and lines representing vector distortions.}
\label{fig:geometricdistortions}
\end{figure}
\subsection{Cosmic Ray Subtraction}
\label{methodology:pipeline:cosmicraysubtraction}
The multiple exposures required for drizzling also allows for the removal of cosmic rays from the images. Cosmic rays can collide with the space-based detectors, leaving characteristic trails across the image (see Figure \ref{fig:cosmicray}). Using observed rates of cosmic rays \cite{WFC3InstrumentHandbook}, we could expect as much as 10\% of an image's pixels to be filled with cosmic rays, given the integration times of our images (see Table \ref{table:observations}). Failure to mask these effects would severely degrade our ability to fit galaxy mophology models and obtain accurate knot fluxes.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=.6\linewidth]{CosmicRay-eps-converted-to.pdf}
\end{center}
\caption{Sample cosmic ray, taken from a F814W image (pre-drizzle). Notice the high signal-to-noise ratio, with a distinctive narrow track.}
\label{fig:cosmicray}
\end{figure}
Fortunately, cosmic rays are transient, and a single ray should not be seen across multiple exposures. By searching for significant differences between exposures, and then masking transients which exhibit the distinctive signature of cosmic rays, the image can be cleaned up substantially. DrizzlePac, in the drizzling process will run a cosmic ray detection and masking algorithm by default, but its parameters should be tweaked as needed (see the DrizzlePac Handbook for more details on the algorithm, and how to ensure that legitimate point sources were not masked \cite{DrizzlePacHandbook}).
\subsection{Error Estimation}
\label{methodology:pipeline:errorestimation}
Finally, after a drizzled image is produced, it is necessary to know the uncertainy associated with the value of each pixel. While the undrizzled exposures have associated error maps (\verb|err| extensions of \verb|*_flt.fits| files), there is no way to produce an exposure-time weighted drizzle with a correct, corresponding uncertainty map \cite{DrizzlePacHandbook}.
In lieu of rewriting a drizzling algorithm which could produce an exposure-time weighted drizzle while simultaneously producing an accurate uncertianty map, we produced estimated error maps. This was accomplished using GALFIT (discussed later in Section \ref{methodology:GALFIT}) which combines in quadrature an estimated read noise (constant across the image; estimated using a median test in a background region) and a Poissonian photon noise \cite{GALFITFAQ}.
\section{Galaxy Subtraction through GALFIT}
\label{methodology:GALFIT}
In order to measure the photometric flux of dim knots against a bright background host galaxy, the GALFIT program \cite{GALFIT} was used to fit and remove standard galactic mophology profiles. At their simplest, these profiles can be Sersic profiles (Eq. \ref{eq:SersicProfile}). For nearby galaxies, it's often necessary to include multiple S\'{e}rsic profiles to adequately remove higher order structure within galaxies (for a discussion on adding multiple components, see \cite{GALFITNearby}), and the presence of an AGN can best be modelled through the inclusion of a point source. To account for instrumental effects, these models were discretized to a grid matching the produced images, and then convolved with the instrument point spread function.
The process of building, convolving and testing a model is then completed for many iterations, searching for best fit model parameters. The user retains control over what models are being fitted, what parameters are free or fixed, and what constraints are to be enforced on the model parameters.
While GALFIT searches for a local minimum using a $\chi^2$ test, there often exist many local minima. After a fit is produced, it is critical that a human evaluates the output residual image, changes the initial values of the free parameters, and continues to reiterate GALFIT as needed.
\subsection{HST Point Spread Functions}
\label{methodology:GALFIT:PSFs}
For both modelling a point-like AGN, and for accounting for diffraction effects inherent to the instrumentation, an accurate point spread function (\emph{PSF}) is necessary. The STScI has created a model PSF creator, TinyTim \cite{TinyTim}, but they note that empirical or extracted PSFs are typically more effective when available \cite{HSTFocus}. In particular, there are a number of known issues applying TinyTim to the WFC3 instrument \cite{TinyTimErrors}. Furthermore, many of the available methods (see \cite{Starfit,TinyTimErrors}) for fitting model PSF parameters (such as the focus of WFC3) are not meant to handle a PSF at the center of an extended, bright galaxy.
Fortunately, an isolated foreground star was captured in the field-of-view of our images. That star was cropped and used as an approximate PSF. That star, and a comparison to the TinyTim predicted PSF, can be seen in Figure \ref{fig:PSF}.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=.8\linewidth]{overlay_contour_3.png}
\end{center}
\caption{Extracted PSF, seen through F160W filter, with TinyTim contours overlaid (note particularly the mismatch of the diffraction spike minima).}
\label{fig:PSF}
\end{figure}
\section{Knot Identification}
\label{methodology:knotidentification}
Jet knots were then identified using the processed files resulting from the HST pipelining (Section \ref{methodology:pipeline}) and the galaxy subtraction (Section \ref{methodology:GALFIT}). Knot candidates were located visually, searching for features which were spatially coincident and similar in morphology to the x-ray and radio jet data (see Figure \ref{fig:XrayRadioMap}). For each knot candidate, windows bounding the candidate were isolated. Inside those windows, 2d-gaussians of specific intensity, $I_\nu$, were fit to the observed data, allowing for the knot's semi-major axis to be misaligned with the jet:
\begin{equation}
I_\nu(x,y) = I_{\nu,0} + I_{\nu,1} \exp\left[ - \frac{ \left(x' - x_0'\right)^2}{2 \sigma_{x'}^2} - \frac{ \left(y' - y_0'\right)^2}{2 \sigma_{y'}^2} \right]
\label{eq:2dgaussian}
\end{equation}
where:
\begin{eqnarray}
x' &=& \left(x- x_0\right) \cos\theta - \left(y - y_0 \right) \sin\theta \nonumber \\
y' &=& \left(x- x_0\right) \sin\theta + \left(y - y_0 \right) \cos\theta \nonumber
\end{eqnarray}
($I_{\nu,0}$, $I_{\nu,1}$, $\sigma_{x'}$, $\sigma_{y'}$, $x_0$, $y_0$, and $\theta$ can all be free parameters in a 2d-gaussian fit.)
The choice of a 2d-gaussian profile is an extension upon a technique previously used on the extragalactic jet PKS-0920 \cite{Marshall01}. In Section \ref{results:knotphotometries} (specifically Table \ref{table:knotcomparisons}) we show that the 2d-gaussian fit reasonably reproduces the total counts that would be obtained by summing the observed counts (preserving photometric acccuracy), while also yeilding a quantitative measure of the solid angle subtended by the source (determined by $\sigma_{x'}$ and $\sigma_{y'}$).
Knot candidates were then flagged if their fitted morphologies did not match that of the x-ray or radio data, if the knot did not fully lie on the detector for all images, or if the fit was poor.
For a sample knot image, with best fit gaussian contours overlaid, see Figure \ref{fig:sample_knot_f160w_32arcsec_intro}.
\section{Knot Photometry}
\label{methodology:knotphotometry}
Photometric fluxes from the knots identified in Section \ref{methodology:knotidentification} were then extracted, and the HST calibration was applied to the observed counts, using for the equation:
\begin{equation}
S_\nu = \frac{DN}{\Delta t} \cdot \mathrm{PHOTFLAM} \cdot \mathrm{PHOTPLAM}^2 \cdot \frac{10^{23}}{2.99 \cdot 10^{18}} \; \mathrm{Jy}
\label{eq:fluxsensetivity}
\end{equation}
for which $DN$ is the number of observed counts (``data numbers''), $\Delta t$ is the total exposure time, $\mathrm{PHOTFLAM}$ and $\mathrm{PHOTPLAM}$ are HST-calibrated constants (respectively: the sensetivity of a given instrument through a given filter, and the pivot wavelength of the filter's passband); the numerical factors are included to give the correct units.
For each window identified in the optical images, the flux over that window was also computed using previously determined x-ray and radio jet profiles (the data reductions were done by Marshall and Lenc for the x-ray and radio data respectively).
\section{Radiative Transfer Modelling}
\label{methodology:radiativetransfer}
Finally, the knot fluxes computed in Section \ref{methodology:knotphotometry} were then used to create numerical spectral models of jet radiative transfer mechanisms. The analytic theory behind these models can be found in Section \ref{theory:radiativetransfer}; the implementation was handled by the code of Krawczynski \cite{Krawczynski2004}.
\chapter{Results}
\label{results}
The theory and methodology of Chapters \ref{theory} and \ref{methodology} was then applied to images taken from the Hubble Space Telescope (see Table \ref{table:observations} for observing details). The results of applying those frameworks to the new optical images, when viewed in light of previous x-ray and radio data, are presented in this chapter.
\section{Galaxy Morphological Components}
\label{results:morphologies}
Using the procedures of Sections \ref{methodology:pipeline} and \ref{methodology:GALFIT}, we cleaned, drizzled and galaxy-subtracted the image. As mentioned previously, the process of fitting galaxy models is degenerate, and there are a number of local $\chi^2_\nu$ minima in the parameter space, requiring significant human guidance in setting the initial fit parameters.
Given that freedom, we focused on minimizing residuals at larger radii, rather than focusing on optimally subtracting the AGN core. This was guided by three insights: the central region would be most susceptible to discrepancies in our PSF, at larger radii any artifacts of the fit would be greatly spread out (appearing as constant offsets at the localized jet knot locations), and the primary x-ray knots (including one which flared) were located at angles greater than 30\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the core.
To see the galaxy-subtracted images, see Figure \ref{fig:galaxysubtractions_largefont}.
\begin{figure}[p]
\centering
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f160w_galaxy_largefont-eps-converted-to.pdf}
\label{fig:galaxysubtractions:f160wgalaxylarge_largefont}
\end{subfigure}
\quad \quad
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f160w_resid_largefont-eps-converted-to.pdf}
\label{fig:galaxysubtractions:f160w_resid}
\end{subfigure} \\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f814w_galaxy_largefont-eps-converted-to.pdf}
\label{fig:galaxysubtractions:f814w_galaxy}
\end{subfigure}
\quad \quad
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f814w_resid_largefont-eps-converted-to.pdf}
\label{fig:galaxysubtractions:f814w_resid}
\end{subfigure}\\
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f475w_galaxy_largefont-eps-converted-to.pdf}
\label{fig:galaxysubtractions:f475w_galaxy}
\end{subfigure}
\quad \quad
\begin{subfigure}{.4\linewidth}
\includegraphics[width=\textwidth]{f475w_resid_largefont-eps-converted-to.pdf}
\label{fig:galaxysubtractions:f475w_resid}
\end{subfigure}
\caption{Images demonstrating the applied galaxy subtractions. (a) images are post-drizzled images, (b) images are the galaxy-subtracted residuals. Note that F160W was taken on a different detector, and for a longer exposure time, leading to the significantly different appearance.}
\label{fig:galaxysubtractions_largefont}
\end{figure}
The best-fit parameters found by GALFIT can be seen in Table \ref{table:morphologies}. For a brief discussion on the family of galaxy models, see Section \ref{theory:morphology}; for a in-depth discussion and motivation for the less standard parameters, see \cite{GALFIT}.
\begin{table*}[p]\centering
\ra{1.3}
\begin{tabular}{lrrrrrr@{}}\toprule
& & {F160W} & \phantom{abc}& {F814W} &
\phantom{abc} & {F475W}\\ \midrule
\emph{PSF} \\
& magnitude, $m$ & 17.5 & & 18.0 & & 17.2 \\
\emph{Sky-subtraction} \\
& ADUs/pixel & -3.0 & & .2 & & .2 \\
\emph{Outer bulge} \\
& magnitude. $m$ & 16.9 & & 15.6 & & 15.5 \\
& S\'{e}rsic index, $n$ & 2.3 & & 2.7 & & 3.1 \\
& effective radius, $R_e$, [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & 16.6 & & 8.4 & & 9.2 \\
& axis ratio, $b/a$ & .6 & & .7 & & .8 \\
& position angle, [\arcdeg] & -31.7 & & -32.0 & & -32.5 \\
& diskyness(-)/boxyness(+) & 0.5 & & & & \\
\emph{Inner bulge} \\
& magnitude. $m$ & 17.4 \\
& S\'{e}rsic index, $n$ & 1.6 \\
& effective radius, $R_e$, [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & 4.2 \\
& axis ratio, $b/a$ & .8 \\
& position angle, [\arcdeg] & -35.0 \\
\bottomrule
\end{tabular}
\caption{\label{table:morphologies} For more details on each parameter, see \cite{GALFIT}. The F160W image, being a deeper image, required an additional S\'{e}rsic bulge to account for the greater range of structure observed. This does not necessarily reflect a different physical morphology.}
\end{table*}
\section{Knot Photometries}
\label{results:knotphotometries}
Using the galaxy-subtracted images (see Figure \ref{fig:galaxysubtractions_largefont}) we looked for features that appeared spatially coincident with the x-ray and radio jets (e.g. Figure \ref{fig:knot_xrayoverlay_f160w_32arcsec}; note that these features are not immediately apparent when viewing the images as a whole). This led us to discover four knot candidates, at 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}, 43\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}, 106\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} and 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the core. (While those angles allow us to compute a project distance from the core, we have little information about their location relative to the core along the line of sight).
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=.8\linewidth]{knot_xrayoverlay_f160w_32arcsec-eps-converted-to.pdf}
\end{center}
\caption{F160W image of the knot located 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the AGN; provided to demonstrate spatial coincidence between x-ray and optical data. Optical data are shown, with x-ray contours overlaid. This is the only knot image with a logarithmic stretch function; all other knot images will have linear stretch functions.}
\label{fig:knot_xrayoverlay_f160w_32arcsec}
\end{figure}
By fitting 2d-gaussians to these knots (see Equation \ref{eq:2dgaussian}) we extracted quantitative measures of knot sizes and fluxes (see Figure \ref{fig:sample_knot_f160w_32arcsec} for an example, Table \ref{table:fittedknots} for reduced photometry results and Appendix \ref{appendix:knotcandidateimages} for all knot candidate images). These knots consistently had the most flux in the F160W images (as opposed to the F814W and F475W images), although in many cases we could only determine upper limits for the F814W and F475W knot candidate fluxes. These fluxes were then checked by simply summing the detected counts across the entire image. This secondary method avoids the systematic biases introduced by assuming a 2d-gaussian profile, and allows for a more direct comparison to the radio and x-ray data (produced by Lenc and Marshall, respectively, who collapsed the transverse data about the jet, leaving only longitudinal information, making a 2d-gaussian fit impossible). As both techniques, 2d-gaussian fitting and simple summing, produce similar results (see Table \ref{table:knotcomparisons}), we feel the data are sufficiently robust. A spectral energy distribution (\emph{SED}) of the observed fluxes can be found in Figure \ref{fig:SED_nomodels_allknots}. (Note: observed fluxes are conventionally reported in Janskys, Jy, such that 1 Jy $ = 10^-23$ erg s$^{-1}$ cm$^{-2}$ Hz$^{-1}$.)
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=.8\linewidth]{f160w_32arcsec-eps-converted-to.pdf}
\end{center}
\caption{F160W image of the knot located 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the AGN; provided to demonstrate the level of agreement between the data and the 2d-gaussian fits. Observed data are shown both in the image and the red contours; green contours show the best fit. The circular white region is a background galaxy which was masked from the analysis.}
\label{fig:sample_knot_f160w_32arcsec}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=.8\linewidth]{SED_no_models_allknots-eps-converted-to.pdf}
\end{center}
\caption{SED of all knot candidates, except the candidate at 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} (for which we are missing data; 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot SED included separately in Figure \ref{fig:SED_nomodels_112arcsec}). Error bars have been omitted, as the error is typically smaller than the markers or the points could only be treated as upper limits for any detectable flux.}
\label{fig:SED_nomodels_allknots}
\end{figure}
One final oddity is the discrepancy between F160W and F814W images of the knot observed at 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} away from the core. While the F160W source is clearly extended, we observe a fully unresolved point source in the F814W images (it has a full width at half maximum, \emph{FWHM}, of .1\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}, approximately equal to the expected PSF FWHM \cite{WFC3InstrumentHandbook}). In many ways this was unexpected: the canonical view of a jet holds that the luminosity profile should not change significantly with small changes in frequency. Furthermore, as a point source, it has a greater intensity but a lower flux than the extended F160W source -- it takes a lower flux and concentrates it within a smaller solid angle on the sky. These results are unusual, and we flag the F814W knot for further analysis. It is worth noting that very little of our analysis would be changed even if chose to remove this point source from our analysis -- Figure \ref{fig:SED_nomodels_only32arcsec} clearly shows a drop in flux between F160W and F475W, whether or not we include the F814W point source.
For the rest of this analysis we will focus on the 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot. This knot corresponds to the brightest x-ray knot \cite{Wilson01}, and was detected with the highest statistical significance (the null hypothesis was ruled out with greater than $5\sigma$ confidence), and does not lie on the edge of any detector chips (unlike the 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot in the F160W image). An SED of just 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot can be seen in Figure \ref{fig:SED_nomodels_only32arcsec}.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=.8\linewidth]{SED_no_models-eps-converted-to.pdf}
\end{center}
\caption{SED of data only from the 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot. Flux uncertainties are smaller than the markers, and thus have been omitted.}
\label{fig:SED_nomodels_only32arcsec}
\end{figure}
\begin{table*}[p]\centering
\ra{1.3}
\begin{tabular}{lrrrrrr}\toprule
& & {F160W} & \phantom{abc}& {F814W} &
\phantom{abc} & {F475W}\\ \midrule
\emph{32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}} \\
& counts & 40000 & & $<$650 & & $<$40 \\
& flux [$\mu$Jy] & 2.2 & & $<$.18 & & $<$.01 \\
& FWHM (x) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & 3.3 & & .02 & & .13 \\
& FWHM (y) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .9 & & .02 & & .02 \\
& rotation [\arcdeg] & 1.1 & & 0.0 & & 25.5 \\
\emph{43\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}} \\
& counts & 6800 & & $<$650 & & $<$740 \\
& flux [$\mu$Jy] & .38 & & $<$.18 & & $<$.11 \\
& FWHM (x) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .30 & & .28 & & .24 \\
& FWHM (y) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .45 & & .17 & & .18 \\
& rotation [\arcdeg] & 51.9 & & -25.0 & & -40.6\\
\emph{106\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}} \\
& counts & 3700 & & $<$360 & & $<$80 \\
& flux [$\mu$Jy] & .21 & & $<$.10 & & $<$.01 \\
& FWHM (x) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .31 & & .19 & & .13 \\
& FWHM (y) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .35 & & .11 & & .09 \\
& rotation [\arcdeg] & 37.7 & & -14.9 & & 0.0 \\
\emph{112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}} \\
& counts & 68000 & & 8200 & & 8300 \\
& flux [$\mu$Jy] & 3.8 & & 2.2 & & 1.2 \\
& FWHM (x) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & 1.2 & & .85 & & 1.2 \\
& FWHM (y) [\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}] & .57 & & .43 & & .42 \\
& rotation [\arcdeg] & -.2 & & 1.2 & & 6.0 \\
\bottomrule
\end{tabular}
\caption{\label{table:fittedknots} Results of 2d-gaussian fits for observed optical knot candidates. The (x) and (y) components of the FWHM correspond to parallel to the jet and perpendicular to the jet, respectively, with a misalignment given by \emph{rotation} (which comes from fitting 2d-gaussians with elliptical contours; (x) and (y) are the semi-major and semi-minor axes). While the 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} knot might have been the strongest detection, it lies on the edge of the F160W detector and was removed from further analyses.}
\end{table*}
\begin{table*}[tbp]\centering
\ra{1.3}
\begin{tabular}{llrrrrr}\toprule
& & {F160W} & \phantom{abc}& {F814W} &
\phantom{abc} & {F475W}\\ \midrule
\emph{32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$} } & \\
flux [$\mu$Jy]& fitted & 2.24 & & $<$.18 & & $<$.01 \\
& summed & 1.98 & & $<$.23 & & $<$.01 \\
\emph{43\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$} } & \\
flux [$\mu$Jy]& fitted & .38 & & $<$.18 & & $<$.11 \\
& summed & .26 & & $<$.17 & & $<$.11 \\
\emph{106\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$} } & \\
flux [$\mu$Jy]& fitted & .21 & & $<$.10 & & $<$.01 \\
& summed & .22 & & $<$.10 & & $<$.01 \\
\emph{112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$} } & \\
flux [$\mu$Jy]& fitted & 3.80 & & 2.24 & & 1.25 \\
& summed & & & 2.47 & & 1.14 \\
\bottomrule
\end{tabular}
\caption{\label{table:knotcomparisons} Comparison of knot flux extraction methods. ``Fitted'' fluxes refer to fitting a 2d-gaussian to the data, and ``summed'' fluxes refer to simply summing the counts within a given window. For more details, see Section \ref{methodology:knotphotometry}. Summed flux is missing from the F160W knot at 112\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}, since that knot extends past the edge of the image.}
\end{table*}
\section{Spectral Energy Distribution Models}
\label{results:seds}
\subsection{Analytic Results}
\label{results:seds:AnalyticResults}
Using the observed knot fluxes seen in Figure \ref{fig:SED_nomodels_only32arcsec} (listed in Table \ref{table:knotcomparisons}), along with data from previous studies \cite{Hardcastle05, Marshall10, Perley97,Tingay08,Wilson01} we can begin to place quantitative constraints on jet emission mechanisms. First, we note that a unbroken power law SED cannot explain all of the observations (see Figure \ref{fig:SED_nomodels_only32arcsec}). The next most commonly applied model is a 2-mechanism model, produced by a single population of electrons (e.g. synchrotron emission at low frequencies, inverse Compton scattering at high frequencies). We will focus mainly on inverse Compton scattering of the cosmic microwave background (\emph{IC-CMB}), simply noting that inverse Compton scattering of synchrotron photons (\emph{SSC}) is numerically shown to be less significant (see Figure \ref{fig:SEDS_p100_p150_p200_p288}).
Assuming a single power law population of electrons (Equation \ref{eq:ElectronDistributionWithCutoffs}) allows us to connect synchrotron observables to inverse Compton observables, placing constraints on both. In particular, theory predicts that the spectral index of synchrotron radiation, $\alpha_{sync}$, be identical to the spectral index for high energy IC-CMB emission, $\alpha_{IC-CMB}$, when far away from the relevant $\gamma_{min}$, $\gamma_{max}$ cutoffs.
While Wilson \cite{Wilson01} and Hardcastle \cite{Hardcastle05} found jet x-ray photon indices of $\Gamma_{x-ray} = 1.94^{+ .43}_{-.49}$ and $\Gamma_{x-ray} = 1.97 \pm .07$ (where $\alpha = \Gamma_{x-ray} - 1$ \cite{XrayGamma} and $S_\nu \sim \nu^{- \alpha}$), those values were determined by combining data for the entire jet. Such an approach would not be appropriate when considering the optical knots -- most of the jet is undetectable at optical frequencies, so we are limited to studying the abnormally luminous regions. We will leave the x-ray spectral index, $\alpha$ (and thus the electron energy index, $p$) unconstrained, and will note the significance of deviations from the results of \cite{Wilson01,Hardcastle05} in Chapter \ref{significance}.
For a given electron index, $p$, we can solve for the predicted values of a number of physical parameters. First, we can constrain $\gamma_{min}$ using location of the low-energy cutoff of IC-CMB emission. Assuming that the IC-CMB SED is above it's low energy cutoff by the time it reaches the frequencies of the observed x-ray emission, we can solve for the smallest $\gamma_{min}$ that can both explain an optical non-detection, and x-ray emission matching our observations (see Table \ref{table:estimatedvalues} for results, and Figure \ref{fig:SED_gamma_min_sample} for an example). Next, we can simultaneously solve for $\gamma_{max}$ and $B_{me}$ using Equations \ref{eq:synchrotronnumaxgammamax} and \ref{eq:BmeWorrall}, using the F160W frequency as $\nu_{max}$ (for the rest of the analysis, we assume $B = B_{me}$ as motivated by \cite{Worrall09}). Knowing $B$ and $\gamma_{max}$ lets us solve for an upper limit on the age of the electron distribution, $t_{age}$ (Equation \ref{eq:electronagelimit}), which we find to be comparable to, but typically greater than the light-crossing time assuming an projected geometry without Doppler boosting. (Electron age limits are a weak constraint, as they do not account for continuing processes; they simply demonstrate physical consistency in the understanding which we are building.) Next, knowing the electron energy cutoffs allows us to find the normalization of the electron energy distribution, $\kappa_e$ which can be determined through Equations \ref{eq:gouldfluxspectrum} and \ref{eq:gouldrhonu}. Finally, those parameters predicts a synchrotron self-absorption frequency, $\nu_{SSA}$, given by Equation \ref{eq:nuSSA}, which must agree to the limits required for a SSA model (with a free $\nu_{SSA}$) to match the radio and optical data.
The results of these calculations, computed for a range of electron indices, $p$, can be found in Table \ref{table:estimatedvalues}.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=.8\linewidth]{gamma_min_sample_p100-eps-converted-to.pdf}
\end{center}
\caption{SED demonstrating the minimal $\gamma_{min}$ constraint, using the IC-CMB cutoff. Modeled IC-CMB emission (red) has been normalized to match the observed x-ray flux, since we are only currently concerned with the spectral shape and location of the cutoff.}
\label{fig:SED_gamma_min_sample}
\end{figure}
\begin{table}[p]\centering
\ra{1.3}
\begin{tabular}{lllrrrrlrrlrrrrrr}\toprule
$p$ & \phantom{a}& $\alpha = \frac{p-1}{2}$ & $\gamma_{min}$ & $\gamma_{max}$ & $B_{me}$ & \phantom{a}& $\kappa_e$ & $\nu_{SSA}$ & \phantom{a}& & $t_{age}$\\
&& & & $10^6$ & $\mu$G && $\mathrm{cm}^{-3}$ & Hz && &$10^3$ yr\\
&& & & & && & \emph{OD} && \emph{SED} \\
\midrule
1.0 && 0.00 & 60 & 2 & 39 && $2 \cdot 10^{-12}$ & $3 \cdot 10^{4}$ && $<5 \cdot 10^{9} $ & 8 \\
1.5 && 0.25 & 80 & 2 & 26 && $4 \cdot 10^{-10}$ & $1 \cdot 10^{5}$ && $<5 \cdot 10^{9} $ & 14 \\
2.001 && 0.5005 & 110 & 3 & 21 && $6 \cdot 10^{-8}$ & $3 \cdot 10^{5}$ && $<5 \cdot 10^{9} $ & 20 \\
2.5 && 0.75 & 130 & 2 & 26 && $5 \cdot 10^{-6}$ & $7 \cdot 10^{5}$ && \phantom{$<$ }$6 \cdot 10^{9}$ & 14 \\
3.0 && 1.00 & 160 & 2 & 37 && $2 \cdot 10^{-4}$ & $1 \cdot 10^{6}$ && \phantom{$<$ }$1 \cdot 10^{10}$ & 8 \\
3.5 && 1.25 & 180 & 2 & 51 && $8 \cdot 10^{-3}$ & $3 \cdot 10^{6}$ && \phantom{$<$ }$3 \cdot 10^{10}$ & 5 \\
4.0 && 1.50 & 200 & 2 & 68 && $3 \cdot 10^{-1}$ & $5 \cdot 10^{6}$ && \phantom{$<$ }$4 \cdot 10^{10}$ & 3 \\
4.5 && 1.75 & 220 & 1 & 85 && $1 \cdot 10^{1}$ & $8 \cdot 10^{6}$ && \phantom{$<$ }$7 \cdot 10^{10}$ & 2 \\
5.0 && 2.00 & 250 & 1 & 103 && $3 \cdot 10^{2}$ & $1 \cdot 10^{7}$ && \phantom{$<$ }$1 \cdot 10^{11}$ & 2 \\
\\
1.90 && 0.45 & 100 & 3 & 21 && $2 \cdot 10^{-8}$ & $2 \cdot 10^{5}$ && \phantom{$<$ }$2 \cdot 10^{9}$ & 19 \\
2.88 && 0.94 & 150 & 2 & 34 && $9 \cdot 10^{-5}$ & $1 \cdot 10^{6}$ && \phantom{$<$ }$1 \cdot 10^{10}$ & 10 \\
3.74 && 1.37 & 190 & 2 & 57 && $5 \cdot 10^{-2}$ & $4 \cdot 10^{6}$ && \phantom{$<$ }$3 \cdot 10^{10}$ & 4 \\
\bottomrule
\end{tabular}
\caption{\label{table:estimatedvalues} Estimates of key physical values, as outline in Section \ref{theory:radiativetransfer}. A representative sample of electron indices, $p$, have been chosen, followed by the electron indices suggested by the data of Wilson \cite{Wilson01} ($p=2.88_{-.90}^{+.86}$, assuming $p = 2 \alpha + 1$). Also note that $p=2$ is a special case not handled by many of the equations used (e.g. Equation \ref{eq:BmeWorrall}), so $p \approx 2$ has been used instead as a representative value. \\ For $\nu_{SSA}$, ``OD'' refers to the estimate predicted by analytic expressions of optical depth (Equation \ref{eq:synchrotronopticaldepth}); ``SED'' denotes the frequency that would be required for a model to match the observed radio and F160W fluxes (the two columns of $\nu_{SSA}$ must match for a self-consistent model). Finally, remember that $t_{age}$ sets only an upper limit on the time since electron re-acceleration, not the actual age of the electron population (see Equation \ref{eq:electronagelimit}).}
\end{table}
\subsection{Numerical Results}
\label{results:seds:NumericalResults}
Using the physical parameters in Table \ref{table:estimatedvalues}, we were able to numerically produce an SED for each electron index, $p$ (see Figure \ref{fig:SEDS_p100_p150_p200_p288}). The first result, which was predicted analytically, is that we need $p \lesssim 2.5$, in order to explain the detection of any knot flux in the F160W image, without requiring the radio flux to be in the self-absorbed regime. This constraint requires that the spectral index for the knot differ significantly from the x-ray indices for the entire jet found by \cite{Hardcastle05,Wilson01}. Further strengthening this limit, we found $p \lesssim 2 $ in order to produce a sharp enough cutoff that could explain a detection in the F160W image, but not in the F814W or F475W images.
Finally, we briefly note that IC-CMB emission always dominates SSC emission, for all choices of $p$ at the x-ray frequency for which we have observations. Neither IC-CMB nor SSC are sufficient for explaining the observed x-ray emission, so for that we must appeal to bulk Doppler boosting, or a different model entirely.
\begin{figure}[p]
\centering
\begin{subfigure}{.49\linewidth}
\includegraphics[width=\textwidth]{p_100_Gamma_1_SED_largefont-eps-converted-to.pdf}
\caption{\label{fig:galaxysubtractions:SEDS_p100} $p=1.00$ }
\end{subfigure}
\begin{subfigure}{.49\linewidth}
\includegraphics[width=\textwidth]{p_150_Gamma_1_SED_largefont-eps-converted-to.pdf}
\caption{\label{fig:galaxysubtractions:SEDS_p150} $p=1.50$ }
\end{subfigure} \\
\begin{subfigure}{.49\linewidth}
\includegraphics[width=\textwidth]{p_200_Gamma_1_SED_largefont-eps-converted-to.pdf}
\caption{\label{fig:galaxysubtractions:SEDS_p200} $p=2.00$ }
\end{subfigure}
\begin{subfigure}{.49\linewidth}
\includegraphics[width=\textwidth]{p_288_Gamma_1_SED_largefont-eps-converted-to.pdf}
\caption{\label{fig:galaxysubtractions:SEDS_p288} $p=2.88$ suggested by \cite{Wilson01} }
\end{subfigure}\\
\caption{Computed SEDs for various electron indices, $p$. Notice that for $p\gtrsim 2.5$ (Figure \ref{fig:galaxysubtractions:SEDS_p288}; the index suggested by the results of \cite{Wilson01,Hardcastle05}), the flux doesn't rise enough after the radio emission to explain the observed F160W flux (without appealing to synchrotron self-absorption, which the data indicate is unlikely). \\
Furthermore for values $p \gtrsim 2$, the synchrotron emission does not drop off quickly enough to explain an F160W detection with an F475W non-detection (\ref{fig:galaxysubtractions:SEDS_p200}). \\
These models also support the prediction that IC-CMB emission (red) would dominate over SSC emission (green). Neither are sufficient though for explaining the x-ray emission, without Doppler boosting or other mechanisms.}
\label{fig:SEDS_p100_p150_p200_p288}
\end{figure}
\subsection{Doppler Boosting}
\label{results:seds:DopplerBoosting}
The discrepancy between the predicted IC-CMB models and the observed x-ray flux suggests the presence of relativistic Doppler boosting. We can use Equations \ref{eq:dopplerfactor} through \ref{eq:MarshallKbetamu}, but we must remember that we cannot uniquely invert Equation \ref{eq:MarshallKbetamu} if both $\Gamma$ and $\theta$ are unknowns. By assuming one, we can solve for the other; by iterating through a number of assumed values, we can determine the extremal values which still predict real values. The results are found in Table \ref{table:dopplerboosting}, for values of $p$ which could explain the radio and optical data ($p \lesssim 2$).
\begin{table}[tbp]\centering
\ra{1.3}
\begin{tabular}{lllrrrr}\toprule
$p$ & \phantom{abc} & $\alpha = \frac{p-1}{2}$ & \phantom{abc} & $\Gamma_{min}$ & \phantom{abc} & $\theta_{max}$ \\
\midrule
1.0 && 0.00 && 233 && 0.1 \\
1.5 && 0.25 && 41 && 0.6 \\
2.001 && 0.5005 && 13 && 2.1 \\
\bottomrule
\end{tabular}
\caption{\label{table:dopplerboosting} Parameters required for Doppler boosting to successfully describe x-ray emission via IC-CMB scattering. $\Gamma$ and $\theta$ are degenerate parameters in our observations, so we can only place limits, by finding the extremal values which yield real values for the other (see Equation \ref{eq:MarshallKbetamu}). Furthermore, note that the inversion does not require $\Gamma_{min}$ to correspond to $\theta_{max}$.}
\end{table}
While the limits in Table \ref{table:dopplerboosting} allow for real values of $\Gamma$ and $\theta$, the systems they predict are unlikely and are not in-line with previous predications about this jet \cite{HSTProposal}. In particular, with $\theta_{max}$ on the order of 1\arcdeg, we would expect the jet, roughly 1 kpc wide, to extend over 5 Mpc from the center of Pictor A, with beaming factors significantly larger than those found in a previous Chandra jet survey \cite{Marshall2005}.
In short, these optical images provide hard constraints on the electron index, $p$, for a successful synchrotron model for the primary knot. Unfortunately, the parameters which could successfully explain the low frequency synchrotron radiation struggle to explain the significant x-ray flux. This tension can be partially alleviated by introducing Doppler boosting, but it is unlikely that Doppler boosting can explain a majority of the high energy flux.
\chapter{Conclusions \& Significance}
\label{significance}
\section{Summarized Status of Radiative Transfer Hypotheses for Pictor A's Jet}
\label{significance:summary}
Early in the study of the jet of Pictor A, data strongly suggested that most of the observed flux was produced by non-thermal processes, such as synchrotron emission and inverse Compton scattering \cite{Wilson01}. The precise nature of this emission was unknown though -- even by the time of Wilson's x-ray data, the results were not strong enough to identify the emission mechanism; in particular, he noted that he could not even rule out an unbroken spectral power law connecting radio and x-ray fluxes.
We have extended the understanding of Pictor A by obtaining optical data and using it to place meaningful constraints on the emission models of Pictor A's jet. To start, our data conclusively rule out an unbroken power law, given the non-detections in the F814W and F475W images (see Figure \ref{fig:SED_nomodels_only32arcsec}); we need different emission mechanisms for different spectral regimes. Spatially, we have shown that the parameters inferred from data of the entire jet differ significantly from parameters inferred from the primary knot, located 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the AGN. This suggests we need a mechanism that explains distinguishably different electron populations at different locations along the jet. Our results suggest that the electron population at the knot 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the AGN core is younger and has a harder spectrum than what we infer for the rest of the jet. Such observations agree with a model of localized electron re-acceleration events (possibly due to shocks, the details of which are poorly understood).
Our data go on to disfavor the canonical model of a single population of electrons emitting non-thermally through synchrotron and inverse Compton processes. We can successfully produce synchrotron models that describe the observed radio and optical data, but none of those models predict enough x-ray flux to match previous observations, even when reasonable levels of Doppler boosting are taken into account.
We cannot definitively determine what is the source of the observed x-ray flux, but we can suggest a direction. Hardcastle originally hypothesized that synchrotron radiation from a second population of electrons could explain the high levels of x-ray flux \cite{Hardcastle05}, a hypothesis motivated by similar results from a different FRII galaxy \cite{Kraft2005}. With the data that were available to Hardcastle, he could not rule out IC-CMB scattering as the primary x-ray producing mechanism, but that appears to have changed with our data. With the addition of our optical data, we have shown it is unlikely for IC-CMB emission to be able to explain the observed x-ray flux. To continue the story of the study of Pictor A,, we suggest specific research in Appendix \ref{appendix:future} which would prove useful for our understanding Pictor A, and by extension, the study of extragalactic jets.
\section*{Acknowledgments}
First and foremost, I would like to thank my thesis adviser, Herman Marshall, for his help and guidance, and for the opportunity to work on this project. I also extend my thanks to the High Energy Astrophysics group at MIT, with whom I have worked for the past year.
For the writing process in particular, I would like to thank Jamie Teherani, the Graduate Resident Tutor of my undergraduate living group, for helping provide a guiding framework with which to approach the writing of this thesis.
For developing my interest in astronomy, I am indebted to professors Paul Schechter and Rob Simcoe, my academic adviser. They helped me find my interest in astronomy, which resulted in this thesis.
Finally, I would like to thank my parents, Abbie and Stuart Gentry, for their support and guidance throughout my studies and the writing of this thesis.
(Support for this work was provided in part by NASA through grant GO-12261.01-A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555.)
\chapter{Introduction}
Some of the most luminous objects of the universe are active galactic nuclei (\emph{AGN}), systems in which the central supermassive blackhole of a galaxy is accreting nearby material, releasing large amounts of energy in the process. In many cases, the luminosity from this central region of a galaxy can exceed the luminosity of the rest of the galaxy combined. This luminosity plays an important role both inside and outside the host galaxy: Within a galaxy, an AGN can play an important role in slowing stellar formation \cite{Olsen:2012me}, and beyond the galaxy itself, AGN also play an important role in bringing about cosmic re-ionization of the intergalactic medium (\emph{IGM}) through photo-ionization. In order to understand the impact of AGN, we must understand what we are observing, and that understanding comes through the study of radiative processes within high energy plasmas.
AGN are not found within every galaxy -- most nearby galaxies, including the Milky Way, currently have dormant supermassive blackholes. It is believed though that most supermassive blackholes, at some point in their lives, become active (astronomers will be watching this year as the gas cloud \emph{G2} approaches our own galactic center, but that accretion event will be much smaller than the accretion onto typical AGN \cite{Gillessen12}). Surveys suggest that AGN were more common at earlier times in the universe, but this poses a challenge to their study: most of the visible AGN are distant (with redshifts, \emph{z}, peaked at $z\approx1$ \cite{Ueda2014}), and are thus fainter and more difficult to resolve. We would like to be able to peer into the structure of AGN, but that is difficult for all but the closest sources.
Pictor A is a particularly interesting galaxy because it is both nearby (redshift, $z$, of .035), and features from its AGN are larger than similar sources. From the AGN core of Pictor A comes a jet of relativistically moving particles, which is about 1 kiloparsec (\emph{kpc}) wide, and extends hundreds of kiloparsecs. For comparison, the nearby galaxy M87 ($z=.004$), has a well-studied jet which is only parsecs wide, and extends only a few kiloparsecs. The combination of Pictor A's size and proximity means we are able to resolve multiple layers of substructure, probing a regime of high energy processes which are not well understood.
In this thesis we will look at recent observations of Pictor A, and make inferences about the radiative processes which are present.
This chapter will look at the background of AGN studies in greater depth, and examine how past research could inform our understanding of Pictor A. In turn, we will also discuss Pictor A, and how it could fill in some of the gaps in our understanding of AGN.
In particular, since AGN often have significant spectral, spatial, and temporal features, this chapter will discuss previous observations of Pictor A in Section \ref{intro:PictorA}. This both provides motivation for the current work, and will factor directly into findings discussed in later chapters. Section \ref{intro:outline} will then give an overview for the research which constitutes the body of this work.
Chapter \ref{theory} treats the subject broadly, giving an overview of the theory guiding this inquiry.
Chapter \ref{methodology} discusses the methodology behind the data collection, processing and analysis, as well as what systemic biases and uncertainties are introduced in our data.
Chapter \ref{results} discusses the observational and reduced results.
Chapter \ref{significance} will consider those reduced parameters in the context of previous observations and physical theories of underlying mechanisms. Discussion of future analyses and recommended observations can be found in Appendix \ref{appendix:future}.
\section{Previous Observations of Pictor A}
\label{intro:PictorA}
Pictor A first attracted attention with its incredibly bright radio emission. Early radio surveys identified it as one of the most luminous radio sources in the sky \cite{Robertson73}. In 1987 it was found that one of its two bright radio lobes was spatially coincident with observed optical synchrotron emission \cite{Roeser87}, and one of the earliest x-ray surveys, the Einstein Imaging Proportional Counter survey (\emph{Einstein IPC}), also partially detected x-ray emission from this lobe \cite{Roeser87}.
Evidence for a jet first came through radio observations, but was most conclusively shown in x-ray images. The radio data came from the Very Large Array (\emph{VLA}), discovering a jet faintly visible extending to the western radio lobe \cite{Perley97}. While faint in the radio, this jet was found to be surprisingly bright in Chandra x-ray observations \cite{Wilson01} (image shown in Fig.\ref{fig:XrayRadioMap}). Follow-up Chandra observations revealed that a particular structure in the jet, a \emph{knot}, appeared to flare and fade on the timescales of years, whereas theory would have na\"{\i}vely predicted a lifetime on the order of thousands of years \cite{Marshall10}. To resolve this, Marshall et al. suggested the flare occurred within an unresolved region within the broader jet (cf. flares observed by Godfrey et al. and Chang et al. \cite{Godfrey09, Chang10}). In order to test this hypothesis, better angular resolution and more spectral information was required.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=.7\linewidth]{0008467v2_fig1_cropped.png}
\end{center}
\caption{Taken from Wilson et al. \cite{Wilson01}. X-ray (Chandra) image overlay \cite{Wilson01}, on top of radio (VLA 20cm) contours \cite{Perley97}. Radio lobes can be seen on either side, but the x-ray jet has only been detected on one side of Pictor A, possibly due to relativistic beaming.}
\label{fig:XrayRadioMap}
\end{figure}
In 2010, Marshall et al. used the Hubble Space Telescope (\emph{HST}) to obtain better spatial resolution and additional spectral information on the nature of the jet. The imaging resolution (around $.05 - .1$ arcseconds, \hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}) is an improvement over Chandra's $.5$\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} resolution, and the chosen filters (with bandpasses centered on wavelengths $1.6\mu$m, $814$nm, $475$nm) were expected to be near a break in the synchrotron spectrum.
In our HST observations, we discovered unresolved substructure, and the photometry revealed a sharp synchrotron cutoff. The significance of these findings will be discussed later.
\section{Outline of this Research}
\label{intro:outline}
This research focuses on 3 HST images of Pictor A, in optical bands ranging from infrared (\emph{IR}) to ultraviolet (\emph{UV}). From these images, we want information about emission coming from relatively atypical jet knots, but those knots are largely drowned out by the large surface brightness of the otherwise typical host galaxy. In order to better identify knot emission, it is then necessary to subtract a model for a ``typical'' elliptical galaxy's surface brightness (see Figure \ref{fig:GalaxySubtraction}), in order to see the atypical features that remain.
\begin{figure}[tbp]
\centering
\begin{subfigure}{.3\linewidth}
\includegraphics[width=\textwidth]{sample_galaxy_subtraction_galaxy-eps-converted-to.pdf}
\label{fig:samplegalaxysubtraction:galaxy}
\caption{Raw image}
\end{subfigure}
\quad
\begin{subfigure}{.3\linewidth}
\includegraphics[width=\textwidth]{sample_galaxy_subtraction_model-eps-converted-to.pdf}
\label{fig:samplegalaxysubtraction:model}
\caption{Model}
\end{subfigure}
\quad
\begin{subfigure}{.3\linewidth}
\includegraphics[width=\textwidth]{sample_galaxy_subtraction_resid-eps-converted-to.pdf}
\label{fig:samplegalaxysubtraction:resid}
\caption{Residual}
\end{subfigure}
\caption{Optical images demonstrating galaxy subtraction. Images are false-color (seen through one band) with a color map and a logarithmic stretch function matching the style of other galaxy-subtracted images (see \cite{GALFITNearby}).}
\label{fig:GalaxySubtraction}
\end{figure}
Using that \emph{galaxy-subtracted} image, we identified features which appeared spatially coincident with observed x-ray and radio features (see knot example, Figure \ref{fig:sample_knot_f160w_32arcsec_intro}) Using these knots, we determined the fluxes through each band. This photometric information was then used to test analytic and numeric models of relativistic synchrotron radiation, synchrotron self-absorption (\emph{SSA}), synchrotron self-Compton radiation (\emph{SSC}), and inverse Compton scattering of the Cosmic Microwave Background (\emph{IC-CMB}).
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=.8\linewidth]{knot_xrayoverlay_f160w_32arcsec-eps-converted-to.pdf}
\end{center}
\caption{Sample knot 32\hbox{$^{\hbox{\rlap{\hbox{\lower4pt\hbox{$\,\prime\prime$}{} from the AGN. Optical image with x-ray contours overlaid. This image is also a false-color image, with a logarithmic stretch function. All future images will also be false-color with logarithmic stretch functions, except when otherwise noted.}
\label{fig:sample_knot_f160w_32arcsec_intro}
\end{figure}
|
\section*{Introduction}
The notion of amenability for monoidal categories first appeared in
Popa's seminal work~\cite{MR1278111} on classification of subfactors
as a crucial condition defining a class of inclusions admitting good
classification. He then gave various characterizations of this property
analogous to the usual amenability conditions for discrete groups: a
Kesten type condition on the norm of the principal graph, a F{\o}lner
type condition on the existence of almost invariant sets, and a
Shannon--McMillan--Breiman type condition on relative entropy, to name
a few.
This stimulated a number of interesting developments in related fields
of operator algebras. First, Longo and Roberts~\cite{MR1444286}
developed a general theory of dimension for C$^*$-tensor categories,
and indicated that the language of sectors/subfactors is well suited
for studying amenability in this context. Then Hiai and
Izumi~\cite{MR1644299} studied amenability for fusion
algebras/hypergroups endowed with a probability measure, and obtained
many characterizations of this property in terms of random walks and
almost invariant vectors in the associated $\ell^p$-spaces. These
studies were followed by the work of Hayashi and
Yamagami~\cite{MR1749868}, who established a way to realize amenable
monoidal categories as bimodule categories over the hyperfinite II$_1$
factor.
In addition to the subfactor theory, another source of interesting
monoidal categories is the theory of quantum groups. In this
framework, the amenability question concerns existence of almost
invariant vectors and invariant means for a discrete quantum group, or
some property of the dimension function on the category of unitary
representations of a compact quantum group~\citelist{\cite{MR1679171}\cite{MR2276175} \cite{MR2113848}}. Here, one should be aware that there are two different notions of amenability involved. One is the coamenability of compact quantum groups (equivalently, amenability of their discrete duals) considered in the regular representations, the
other is the amenability of representation categories. These notions
coincide only for quantum groups of Kac type.
\smallskip
In yet another direction, Izumi~\cite{MR1916370} developed a theory of
noncommutative Poisson boundary for discrete quantum groups in order to study
the minimality (or lack thereof) of infinite tensor product type
actions of compact quantum groups. From the subsequent
work~\citelist{\cite{MR2200270}\cite{MR2335776}} it became
increasingly clear that for coamenable compact quantum groups the
Poisson boundary captures a very elaborate difference between the two
amenability conditions. Later, an important result on noncommutative
Poisson boundaries was obtained by De Rijdt and Vander
Vennet~\cite{MR2664313}, who found a way to compute the boundaries
through monoidal equivalences. In light of the categorical duality
for compact quantum group actions recently developed
in~\citelist{\cite{MR3121622}\cite{neshveyev-mjm-categorification}},
this result suggests that the Poisson boundary should really be an
intrinsic notion of the representation category $\Rep G$ itself,
rather than of the choice of a fiber functor giving a concrete
realization of $\Rep G$ as a category of Hilbert spaces. Starting from
this observation, in this paper we define Poisson boundaries for
monoidal categories.
To be more precise, our construction takes a rigid C$^*$-tensor
category $\mathcal{C}$ with simple unit and a probability measure $\mu$ on the
set $\Irr(\mathcal{C})$ of isomorphism classes of simple objects, and gives
another C$^*$-tensor category $\mathcal{P}$ together with a unitary tensor
functor $\Pi\colon\mathcal{C}\to\mathcal{P}$. Although the category~$\mathcal{P}$ is defined
purely categorically, there are several equivalent ways to describe
it, or at least its morphism sets, that are more familiar to the operator
algebraists. One is an analogue of the standard description of
classical Poisson boundaries as ergodic components of the time
shift. Another is in terms of relative commutants of von Neumann
algebras, in the spirit
of~\citelist{\cite{MR1444286}\cite{MR1749868}\cite{MR1916370}}. For
categories arising from subfactors and quantum groups, this can be
made even more concrete. For subfactors, computing the Poisson boundary essentially corresponds to passing to the standard model of a subfactor~\cite{MR1278111}. For quantum groups, not surprisingly as this was our initial motivation, the Poisson boundary of the representation category of $G$ can be described in terms of the Poisson boundary of~$\hat G$. The last
result will be discussed in detail in a separate
publication~\cite{arXiv:1310.4407}, since we also want to describe the
action of $\hat G$ on the boundary in categorical terms and this would
lead us away from the main subject of this paper.
\smallskip
Our main result is that if $\mathcal{P}$ has simple unit, which corresponds to
ergodicity of the classical random walk defined by $\mu$ on
$\Irr(\mathcal{C})$, then $\Pi\colon\mathcal{C}\to\mathcal{P}$ is a universal unitary tensor
functor which induces the amenable dimension function on $\mathcal{C}$. From
this we conclude that $\mathcal{C}$ is amenable if and only if there exists a
measure $\mu$ such that $\Pi$ is a monoidal equivalence. The last
result is a direct generalization of the famous characterization of
amenability of discrete groups in terms of their Poisson boundaries
due to Furstenberg~\cite{MR0352328}, Kaimanovich and
Vershik~\cite{MR704539}, and Rosenblatt~\cite{MR630645}. From this
comparison it should be clear that, contrary to the usual
considerations in subfactor theory, it is not enough to work only with
finitely supported measures, since there are amenable groups which do
not admit any finitely supported ergodic
measures~~\cite{MR704539}. The characterization of amenability in
terms of Poisson boundaries generalizes several results
in~\citelist{\cite{MR1278111} \cite{MR1444286}\cite{MR1749868}}. Our
main result also allows us to describe functors that factor through
$\Pi$ in terms of categorical invariant means. For quantum groups
this essentially reduces to the equivalence between coamenability of
$G$ and amenability of~$\hat
G$~\citelist{\cite{MR2276175}\cite{MR2113848}}.
\smallskip
Although our theory gives a satisfactory unification of various
amenability results, the main remarkable property of the functor
$\Pi\colon\mathcal{C}\to\mathcal{P}$ is, in our opinion, the universality. If the
category $\mathcal{P}$ happens to have a simpler structure compared to $\mathcal{C}$,
this universality allows one to reduce classification of functors from
$\mathcal{C}$ inducing the amenable dimension function, to an easier
classification problem for functors from $\mathcal{P}$. This idea will be used
in~\cite{classification} to classify a class of compact quantum
groups.
\smallskip
\paragraph{\bf Acknowledgement} M.Y.~thanks M.~Izumi, S.~Yamagami, T.~Hayashi, and R.~Tomatsu for their interest and encouragement at various stages of
the project.
\bigskip
\section{Preliminaries}
\label{sec:preliminaries}
\subsection{Monoidal categories}
The principal subject of this paper is \emph{rigid C$^*$-tensor
categories}. By now there are many materials covering the basics of
this subject, see for
example~\citelist{\cite{MR2091457}\cite{arXiv:0804.3587v3} \cite{neshveyev-tuset-book}}
and references therein. We mainly follow the conventions
of~\cite{neshveyev-tuset-book}, but for the convenience of the reader
we summarize the basic definitions and facts below.
\smallskip
A \emph{C$^*$-category} is a category $\mathcal{C}$ whose morphism sets
$\mathcal{C}(U, V)$ are complex Banach spaces endowed with complex conjugate
involution $\mathcal{C}(U, V) \to \mathcal{C}(V, U)$, $T \mapsto T^*$ satisfying the
C$^*$-identity. Unless said otherwise, we always assume that $\mathcal{C}$ is
closed under direct sums and subobjects. The latter means that any
idempotent in the endomorphism ring $\End{\mathcal{C}}{X} = \mathcal{C}(X, X)$ comes
from a direct summand of~$X$.
A C$^*$-category is said to be \emph{semisimple} if any object is
isomorphic to a direct sum of simple (that is, with the endomorphism
ring $\mathbb{C}$) objects. We then denote the isomorphism classes of simple
objects by~$\Irr (\mathcal{C})$ and assume that this is an at most countable
set. Many results admit formulations which do not require this
assumption and can be proved by considering subcategories generated by
countable sets of simple objects, but we leave this matter to the
interested reader.
A \emph{unitary functor}, or a \emph{C$^*$-functor}, is a linear
functor of C$^*$-categories $F\colon \mathcal{C} \to \mathcal{C}'$ satisfying $F(T^*)
= F(T)^*$.
In this paper we frequently perform the following operation: starting
from a C$^*$-category $\mathcal{C}$, we replace the morphisms sets by some
larger system $\mathcal{D}(X, Y)$ naturally containing the original $\mathcal{C}(X,
Y)$. Then we perform the \emph{idempotent completion} to construct a
new category $\mathcal{D}$. That is, we regard the projections $p
\in \End{\mathcal{D}}{X}$ as objects in the new category, and take $q \mathcal{D}(X,
Y) p$ as the morphism set from the object represented by $p
\in \End{\mathcal{D}}{X}$ to the one by $q \in \End{\mathcal{D}}{Y}$. Then the
embeddings $\mathcal{C}(X, Y) \to \mathcal{D}(X, Y)$ can be considered as a
C$^*$-functor $\mathcal{C} \to \mathcal{D}$.
A \emph{C$^*$-tensor category} is a C$^*$-category endowed with a
unitary bifunctor $\otimes \colon \mathcal{C} \times \mathcal{C} \to \mathcal{C}$, a
distinguished object $\mathds{1} \in \mathcal{C}$, and natural isomorphisms
\begin{align*} \mathds{1} \otimes U &\simeq U \simeq U \otimes \mathds{1},& \Phi({U,
V, W})&\colon (U \otimes V) \otimes W \to U \otimes (V \otimes W)
\end{align*} satisfying certain of compatibility conditions.
A \emph{unitary tensor functor}, or a \emph{C$^*$-tensor functor},
between two C$^*$-tensor categories $\mathcal{C}$ and $\mathcal{C}'$ is given by a
triple $(F_0, F, F_2)$, where $F$ is a C$^*$-functor $\mathcal{C} \to \mathcal{C}'$,
$F_0$ is a unitary isomorphism $\mathds{1}_{\mathcal{C}'} \to F(\mathds{1}_\mathcal{C})$, and $F_2$
is a natural unitary isomorphism $F(U) \otimes F(V) \to F(U \otimes
V)$, which are compatible with the structure morphisms of $\mathcal{C}$ and
$\mathcal{C}'$. As a rule, we denote tensor functors by just one symbol
$F$.
When $\mathcal{C}$ is a strict C$^*$-tensor category and $U \in \mathcal{C}$, an
object $V$ is said to be a \emph{dual object} of $U$ if there are
morphisms $R \in \mathcal{C}(\mathds{1}, V \otimes U)$ and $\bar{R} \in \mathcal{C}(\mathds{1}, U
\otimes V)$ satisfying the conjugate equations
\begin{align*} (\iota_{V} \otimes \bar{R}^*) (R \otimes \iota_{V}) &=
\iota_{V},& (\iota_{U} \otimes R^*) (\bar{R} \otimes \iota_{U}) &=
\iota_{U}.
\end{align*} If any object in $\mathcal{C}$ admits a dual, $\mathcal{C}$ is said to be
\emph{rigid} and we denote a choice of a dual of $U \in \mathcal{C}$
by~$\bar{U}$. A rigid C$^*$-tensor category (with simple unit) has
finite dimensional morphism spaces and hence is automatically
semisimple. The quantity
$$
d^\mathcal{C}(U) =\min_{(R, \bar{R})} \norm{R} \norm{\bar{R}}
$$
is called the \emph{intrinsic dimension} of $U$, where $(R, \bar{R})$
runs through the set of solutions of conjugate equations as above. We
omit the superscript $\mathcal{C}$ when this is no danger of confusion. A
solution $(R, \bar{R})$ of the conjugate equations for $U$ is called
\emph{standard} if
$$
\|R\|=\|\bar R\|=d(U)^{1/2}.
$$
Solutions of the conjugate equations for $U$ are unique up to the
transformations $$(R, \bar{R}) \mapsto ((T^* \otimes \iota) R, (\iota
\otimes T^{-1}) \bar{R}).$$ Furthermore, if $(R,\bar R)$ is standard,
then such a transformation defines a standard solution if and only if
$T$ is unitary.
In a rigid C$^*$-tensor category $\mathcal{C}$ we often fix standard solutions
$(R_U,\bar R_U)$ of the conjugate equations for every object $U$. Then
$\mathcal{C}$ becomes \emph{spherical} in the sense that one has the equality
$R_U^* (\iota \otimes T) R_U = \bar{R}_U^* (T \otimes \iota)
\bar{R}_U$ for any $T \in \End{\mathcal{C}}{U}$. The normalized linear
functional
$$
\tr_U(T) = \frac{1}{d(U)} R_U^* (\iota \otimes T)
R_U=\frac{1}{d(U)}\bar{R}_U^* (T \otimes \iota) \bar{R}_U
$$
is a tracial state on the finite dimensional C$^*$-algebra
$\End{\mathcal{C}}{U}$. It is independent of the choice of a standard
solution.
Given a rigid C$^*$-tensor category $\mathcal{C}$, if $[U]$ and $[V]$ are
elements of $\Irr(\mathcal{C})$, we can define their product in
$\mathbb{Z}_+[\Irr(\mathcal{C})]$ by putting
$$
[U] \cdot [V] = \sum_{[W] \in \Irr (\mathcal{C})} \dim \mathcal{C}(W, U \otimes V)
[W],
$$
thus getting a semiring $\mathbb{Z}_+[\Irr(\mathcal{C})]$. Extending this formula by
bilinearity, we obtain a ring structure on $\mathbb{Z}[\Irr(\mathcal{C})]$. The map
$[U] \mapsto d(U)$ extends to a ring homomorphism $\mathbb{Z}[\Irr(\mathcal{C})] \to
\mathbb{R}$. The pair $(\mathbb{Z}[\Irr(\mathcal{C})], d)$ is called the \emph{fusion
algebra} of $\mathcal{C}$. In general, a ring homomorphism $d'\colon
\mathbb{Z}[\Irr(\mathcal{C})] \to \mathbb{R}$ satisfying $d'([U]) > 0$ and $d'([U])=d'([\bar
U])$ for every $[U] \in \Irr(\mathcal{C})$ is said to be a \emph{dimension
function} on $\mathcal{C}$.
For a rigid C$^*$-tensor category $\mathcal{C}$, the right multiplication by
$[U] \in \Irr(\mathcal{C})$ on $\mathbb{Z}[\Irr(\mathcal{C})]$ can be considered as a densely
defined operator $\Gamma_U$ on $\ell^2(\Irr(\mathcal{C}))$. This definition
extends to arbitrary objects of $\mathcal{C}$ by the formula $\Gamma_U =
\sum_{[V] \in \Irr(\mathcal{C})} \dim(V, U) \Gamma_V$. If $d'$ is a dimension
function on $\mathcal{C}$, one has the estimate $$\norm{\Gamma_U}_{B(\ell^2(
\Irr(\mathcal{C})))} \le d'(U).$$ If the equality holds for all objects $U$,
then the dimension function $d'$ is called \emph{amenable}. Clearly,
there can be at most one amenable dimension function. If the intrinsic
dimension function is amenable, then $\mathcal{C}$ itself is called amenable.
\subsection{Categories of functors} \label{sstensorfunctors}
Given a rigid C$^*$-tensor category $\mathcal{C}$ we will consider the
category of unitary tensor functors from $\mathcal{C}$ into C$^*$-tensor
categories. Its objects are pairs $(\mathcal{A},E)$, where $\mathcal{A}$ is a
C$^*$-tensor category and $E\colon\mathcal{C}\to\mathcal{A}$ is a unitary tensor
functor. The morphisms $(\mathcal{A},E)\to(\mathcal{B},F)$ are unitary tensor functors
$G\colon\mathcal{A}\to\mathcal{B}$, considered up to natural unitary monoidal
isomorphisms,\footnote{Therefore the category of functors from $\mathcal{C}$
we consider here is different from the category $\mathpzc{Tens}(\mathcal{C})$ defined
in \cite{arXiv:1310.4407}, where we wanted to distinguish between
isomorphic functors and defined a more refined notion of morphisms.}
such that $GE$ is naturally unitarily isomorphic to $F$.
A more concrete way of thinking of this category is as follows. First
of all we may assume that~$\mathcal{C}$ is strict. Consider a unitary tensor
functor $E\colon\mathcal{C}\to\mathcal{A}$. The functor $E$ is automatically faithful
by semisimplicity and existence of conjugates in $\mathcal{C}$. It follows
that by replacing the pair $(\mathcal{A},E)$ by an isomorphic one, we may
assume that $\mathcal{A}$ is a strict C$^*$-tensor category containing $\mathcal{C}$
and $E$ is simply the embedding functor. Namely, define the new sets
of morphisms between objects $U$ and $V$ in $\mathcal{C}$ as $\mathcal{A}(E(U),E(V))$,
and then complete the category we thus obtain with respect to
subobjects.
Assume now that we have two strict C$^*$-tensor categories $\mathcal{A}$ and
$\mathcal{B}$ containing $\mathcal{C}$, and let $E\colon\mathcal{C}\to \mathcal{A}$ and
$F\colon\mathcal{C}\to\mathcal{B}$ be the embedding functors. Assume $[G]\colon
(\mathcal{A},E)\to(\mathcal{B},F)$ is a morphism. This means that there exist unitary
isomorphisms $\eta_U\colon G(U)\to U$ in $\mathcal{B}$ such that
$G(T)=\eta_V^{-1}T\eta_U$ for any morphism $T\in\mathcal{C}(U,V)$, and the
morphisms $$G_{2}(U,V)\colon G(U)\otimes G(V)\to G(U\otimes V)$$
defining the tensor structure of $G$ restricted to $\mathcal{C}$ are given by
$G_{2}(U,V)=\eta_{U\otimes V}^{-1}(\eta_U\otimes\eta_V)$. For
objects~$U$ of~$\mathcal{A}$ that are not in $\mathcal{C}$ put
$\eta_U=1\in\mathcal{B}(G(U))$. We can then define a new unitary tensor
functor $\tilde G\colon A\to\mathcal{B}$ by letting $\tilde G(U)=U$ for
objects $U$ in $\mathcal{C}$ and $\tilde G(U)=G(U)$ for the remaining objects,
$\tilde G(T)=\eta_VG(T)\eta_U^{-1}$ for morphisms, and $\tilde
G_2(U,V)=\eta_{U\otimes
V}G_2(U,V)(\eta_U^{-1}\otimes\eta_V^{-1})$. Then $[G]=[\tilde G]$ and
the restriction of $\tilde G$ to $\mathcal{C}\subset\mathcal{A}$ coincides with the
embedding (tensor) functor $\mathcal{C}\to\mathcal{B}$.
Therefore, any unitary tensor functor $\mathcal{C}\to\mathcal{A}$ is naturally unitary
isomorphic to an embedding functor, and the morphisms between two such
embeddings $E\colon\mathcal{C}\to\mathcal{A}$ and $F\colon\mathcal{C}\to\mathcal{B}$ are the unitary
tensor functors $G\colon\mathcal{A}\to\mathcal{B}$ extending $F$, considered up to
natural unitary isomorphisms. If, furthermore, $\mathcal{A}$ is generated by
the objects of $\mathcal{C}$ then $[G]$ is completely determined by the maps
$\mathcal{A}(U,V)\to\mathcal{B}(U,V)$ extending the identity maps on $\mathcal{C}(U,V)$ for all
objects $U$ and $V$ in $\mathcal{C}$.
\subsection{Subfactor theory}
Let $N \subset M$ be an inclusion of von Neumann algebras represented
on a Hilbert space $H$. There is a canonical bijective correspondence
between the normal semifinite faithful operator valued weights $\Phi
\colon M \to N$ and the ones $\Psi\colon N' \to M'$ in terms of
spatial derivatives~\cite{MR561983}. Given such a $\Phi$, one
associates a unique $\Psi$ denoted by $\Phi^{-1}$ and characterized by
the equation
$$
\frac{d \omega \Phi}{d \omega'} = \frac{d \omega}{d \omega'
\Phi^{-1}},
$$
where $\omega$ and $\omega'$ are any choices of normal semifinite
faithful weights on $N$ and $M'$.
If $E$ is a normal faithful conditional expectation from $M$ to $N$,
its \emph{index} $\Ind E$ can be defined as
$E^{-1}(1)$~\cite{MR829381}. Suppose that $M$ and $N$ are factors
admitting conditional expectations of finite index. Then
there is a unique choice of $E$ which minimizes $\Ind E$. This $E$ is
called the \emph{minimal conditional expectation} of the subfactor $N
\subset M$~\cite{MR976765}.
Suppose that $N \subset M$ is a subfactor endowed with a normal
conditional expectation of finite index $E\colon M \to N$. We then
obtain a von Neumann algebra $M_1$ called the \emph{basic extension}
of $N\subset M$ with respect to $E$, as follows. Taking a normal
semifinite faithful weight $\psi$ on $N$, the algebra $M_1 \subset
B(L^2(M, \psi E))$ is generated by~$M$ and the orthogonal projection
$e_N$, called the Jones projection, onto $L^2(N, \psi) \subset L^2(M,
\psi E)$. One has the equality $M_1 = J N' J$, where $J$ is the
modular conjugation of $M$ with respect to~$\psi E$. From the above
correspondence of operator valued weights, there is a canonical
conditional expectation $E_1 \colon M_1 \to M$ which has the same
index as $E$, namely, $E_1=(\Ind E)^{-1}JE^{-1}(J\cdot
J)J$. Iterating this procedure, we obtain a tower of von Neumann
algebras
$$
N \subset M \subset M_1 \subset M_2 \subset \cdots.
$$
The higher relative commutants
$$
N' \cap M_k = \{ x \in M_k \mid \forall y \in N \colon x y = y x \}
$$
are finite dimensional C$^*$-algebras. The algebras $M' \cap M_{2 k}$
($k \in \mathbb{N}$) can be considered as the endomorphism rings of $M
\otimes_N M \otimes_N \cdots \otimes_N M$ in the category of
$M$-bimodules, and there are similar interpretations for the $N' \cap
M_{2k + 1}$, etc., in terms of the $N$-bimodules, $M$-$N$-modules, and
$N$-$M$-modules.
\subsection{Relative entropy}
An important numerical invariant for inclusions of von Neumann
algebras, closely related to index, is relative entropy. For this part
we follow the exposition of~\cite{MR2251116}.
When $\varphi$ and
$\psi$ are positive linear functionals on a C$^*$-algebra $M$, we
denote their relative entropy by $S(\varphi,\psi)$. If $M$ is finite
dimensional, it can be defined as
$$
S(\varphi,\psi)=\begin{cases}\Tr(Q_\varphi(\log Q_\varphi-\log
Q_\psi)),& \text{if}\ \ \varphi\le\lambda\psi\ \ \text{for some}\ \
\lambda>0,\\+\infty,&\text{otherwise,}\end{cases}
$$
where $\Tr$ is the canonical trace on $M$, which takes value $1$ on
every minimal projection in $M$, and $Q_\varphi\in M$ is the density
matrix of $\varphi$, so that we have $\varphi(x) =\Tr( x
Q_\varphi)$. For a single positive linear functional $\psi$ on a
finite dimensional $M$, we also have its von Neumann entropy defined
as $S(\psi)=-\Tr(Q_\psi\log Q_\psi)$.
\smallskip
Given an inclusion of C$^*$-algebras $N\subset M$ and a state
$\varphi$ on $M$, the \emph{relative entropy} $H_\varphi(M|N)$ (also
called \emph{conditional entropy} in the classical probability
theory) is defined as the supremum of the quantities
$$
\sum_i(S(\varphi_i,\varphi)-S(\varphi_i|_N,\varphi|_N))
$$
where $(\varphi_i)_i = (\varphi_1, \ldots, \varphi_k)$ runs through
the tuples of positive linear functionals on $M$ satisfying $\varphi =
\sum^k_{i=1} \varphi_i$. If $M$ is finite dimensional, this can also
be written as
$$
H_\varphi(M | N) = S(\varphi) - S(\varphi|_N) + \sup_{(\varphi_i)_i}
\sum_i \bigl(S(\varphi_i|_N) - S(\varphi_i)\bigr),
$$
where supremum is again taken over all finite decompositions of
$\varphi$.
Relative entropy has the following lower semicontinuity
property. Suppose that $N \subset M$ is an inclusion of von Neumann
algebras and $\varphi$ is a normal state on $M$. Suppose that $B_i
\subset A_i$ ($i=1,2,\ldots$) are increasing sequences of subalgebras
$B_i \subset N$, $A_i \subset M$ such that $\cup_i A_i$ and $\cup_i
B_i$ are $s^*$-dense in~$M$ and~$N$, respectively. Then one has
the estimate $$H_\varphi(M | N) \le \liminf_i H_\varphi(A_i | B_i).$$
If $N \subset M$ is an inclusion of von Neumann algebras and $E \colon
M \to N$ is a normal conditional expectation, the relative entropy of
$M$ and $N$ with respect to $E$ is defined by
$$
H_E(M | N) = \sup_\varphi H_\varphi(M | N),
$$
where $\varphi$ runs through the normal states on $M$ satisfying
$\varphi = \varphi E$~\cite{MR1096438}. If $M$ and $N$ are factors,
then we have the estimate $H_E(M|N)\le\log\Ind E$.
\bigskip
\section{Categorical Poisson boundary}
Let $\mathcal{C}$ be a strict rigid C$^*$-tensor category satisfying our
standard assumptions: it is closed under direct sums and subobjects,
the tensor unit is simple, and $\Irr(\mathcal{C})$ is at most countable.
\smallskip
Let $\mu$ be a probability measure on $\Irr(\mathcal{C})$. The Poisson
boundary of $(\mathcal{C},\mu)$ will be a new C$^*$-tensor category $\mathcal{P}$,
possibly with nonsimple unit, together with a unitary tensor functor
$\Pi\colon\mathcal{C}\to\mathcal{P}$. In this section we define $(\mathcal{P},\Pi)$ in purely
categorical terms. In the next section we will give several more
concrete descriptions of this construction.
\smallskip
Recall that $\tr_X$ stands for the normalized categorical
trace on $\End{\mathcal{C}}{X}$. More generally, for objects~$U$ and~$V$ we
denote by
$$
\tr_X\otimes\iota\colon\mathcal{C}(X\otimes U,X\otimes V)\to\mathcal{C}(U,V)
\ \ \text{and}\ \ \iota\otimes\tr_X\colon\mathcal{C}(U\otimes X,V\otimes
X)\to\mathcal{C}(U,V)
$$
the normalized partial categorical traces. Namely, if
$R_X\colon\mathds{1}\to\bar X\otimes X$ and $\bar R_X\colon \mathds{1}\to
X\otimes\bar X$ form a standard solution of the conjugate equations
for $X$, then
$$
(\tr_X\otimes\iota)(T)=d(X)^{-1}(R_X^*\otimes\iota)(\iota\otimes
T)(R_X\otimes\iota),
\ \ (\iota\otimes\tr_X)(T)=d(X)^{-1}(\iota\otimes\bar
R_X^*)(T\otimes\iota)(\iota\otimes\bar R_X).
$$
For an object $U$ consider the functor $\iota\otimes
U\colon\mathcal{C}\to\mathcal{C}$. Given two objects $U$ and $V$, consider the space
$\Nat(\iota\otimes U,\iota\otimes V)$ of natural transformations from
$\iota\otimes U$ to $\iota\otimes V$, so elements of
$\Nat(\iota\otimes U,\iota\otimes V)$ are collections
$\eta=(\eta_X)_X$ of natural in $X$ morphisms $\eta_X\colon X\otimes
U\to X\otimes V$. For every object $X$ we can define a linear operator
$P_X$ on $\Nat(\iota\otimes U,\iota\otimes V)$ by
$$
P_X(\eta)_Y=(\tr_X\otimes\iota)(\eta_{X\otimes Y}).
$$
Denote by $\hat{\mathcal{C}}(U, V) \subset \Nat(\iota\otimes U,\iota\otimes
V)$ the subspace of bounded natural transformations, that is, of
elements $\eta$ such that $\sup_Y\|\eta_Y\|<\infty$. This is a Banach
space, and the operator $P_X$ defines a contraction on it. It is also
clear that the operator $P_X$ depends only on the isomorphism class of
$X$.
For every $s\in \Irr(\mathcal{C})$ fix a representative $U_s$. We write
$\tr_s$ instead of $\tr_{U_s}$, $P_s$ instead of $P_{U_s}$, and so
on. Similarly, for a natural transformation $\eta\colon\iota\otimes
U\to\iota\otimes V$ we will write $\eta_s$ instead of
$\eta_{U_s}$. Let also denote by $e\in\Irr(\mathcal{C})$ the index
corresponding to $\mathds{1}$. For convenience we assume that
$U_e=\mathds{1}$. Define an involution on $\Irr(\mathcal{C})$ such that $U_{\bar s}$
is a dual object to $U_s$.
Consider now the operator
$$
P_\mu=\sum_s\mu(s)P_s.
$$
This is a well-defined contraction on $\hat{\mathcal{C}}(U, V)$. We say that a
bounded natural transformation $\eta\colon\iota\otimes
U\to\iota\otimes V$ is $P_\mu$-{\em harmonic} if
$$P_\mu(\eta)=\eta.$$
Any morphism $T\colon U\to V$ defines a bounded natural transformation
$(\iota_X\otimes T)_X$, which is obviously $P_\mu$-harmonic for every
$\mu$. When there is no ambiguity we denote this natural
transformation simply by $T$.
The composition of harmonic transformations is in general not
harmonic. But we can define a new composition as follows.
\begin{proposition} \label{pproduct} Given bounded $P_\mu$-harmonic
natural transformations $\eta\colon\iota\otimes U\to\iota\otimes V$
and $\nu\colon\iota\otimes V\to\iota\otimes W$, the limit
$$
(\nu\cdot\eta)_X=\lim_{n\to\infty}P^n_\mu(\nu\eta)_X
$$
exists for all objects $X$ and defines a bounded $P_\mu$-harmonic
natural transformation $\iota\otimes U\to\iota\otimes W$. Furthermore,
the composition $\cdot$ is associative.
\end{proposition}
Note that since the spaces $\mathcal{C}(X\otimes U,X\otimes W)$ are finite
dimensional by our assumptions on $\mathcal{C}$, the notion of a limit is
unambiguous.
\begin{proof}[Proof of Proposition~\ref{pproduct}] This is an immediate
consequence of results of Izumi~\cite{MR2995726} (another proof will
be given in Section~\ref{sec:poiss-bound-from}). Namely, replacing
$U$, $V$ and $W$ by their direct sum we may assume that
$U=V=W$. Then
$$
\hat{\mathcal{C}}(U) = \hat{\mathcal{C}}(U, U) \cong
\ell^\infty\text{-}\bigoplus_s\End{\mathcal{C}}{U_s\otimes U}
$$
is a von Neumann algebra and $P_\mu$ is a normal unital completely
positive map on it. By \cite{MR2995726}*{Corollary~5.2} the subspace
of $P_\mu$-invariant elements is itself a von Neumann algebra with
product $\cdot$ such that $x\cdot y$ is the $s^*$-limit of the
sequence $\{P^n_\mu(xy)\}_n$. \end{proof}
Using this product on harmonic elements we can define a new
C$^*$-tensor category $\mathcal{P}=\mathcal{P}_{\mathcal{C},\mu}$ and a unitary tensor functor
$\Pi=\Pi_{\mathcal{C},\mu}\colon\mathcal{C}\to\mathcal{P}$ as follows.
First consider the category $\tilde\mathcal{P}$ with the same objects as in
$\mathcal{C}$, but define the new spaces $\tilde\mathcal{P}(U,V)$ of morphisms as the
spaces of bounded $P_\mu$-harmonic natural transformations
$\iota\otimes U\to\iota\otimes V$. Define the composition of morphisms
as in Proposition~\ref{pproduct}. We thus get a C$^*$-category,
possibly without subobjects. Furthermore, the C$^*$-algebras
$\End{\tilde\mathcal{P}}{U}$ are von Neumann algebras.
Next, we define the tensor product of objects in the same way as in
$\mathcal{C}$, and define the tensor product of morphisms by
$$
\nu\otimes\eta=(\nu\otimes\iota)\cdot(\iota\otimes\eta).
$$
Here, given $\nu\colon \iota\otimes U\to\iota\otimes V$ and
$\eta\colon\iota\otimes W\to\iota\otimes Z$, the natural
transformation $\nu\otimes\iota_Z\colon\iota\otimes U\otimes
Z\to\iota\otimes V\otimes Z$ is defined by
$$
(\nu\otimes\iota_Z)_X=\nu_X\otimes\iota_Z,
$$
while the natural transformation $\iota_U\otimes\eta\colon\iota\otimes
U\otimes W\to\iota\otimes U\otimes Z$ is defined by
$$
(\iota_U\otimes\eta)_X=\eta_{X\otimes U}.
$$
We remark that $\nu\otimes\iota$ and $\iota\otimes\eta$ are still
$P_\mu$-harmonic due to the identities
$$
P_X(\nu\otimes\iota) = P_X(\nu)\otimes\iota,\ \ P_X(\iota\otimes\eta)
= \iota\otimes P_X(\eta).
$$
Note also that by naturality of $\eta$ we have
$(\nu_X\otimes\iota_Z)\eta_{X\otimes U}=\eta_{X\otimes
V}(\nu_X\otimes\iota_Z)$, which implies that
$$
\nu\otimes\eta=(\iota\otimes\eta)\cdot(\nu\otimes\iota).
$$
This shows that $\otimes\colon\tilde\mathcal{P}\times\tilde\mathcal{P}\to\tilde\mathcal{P}$ is
indeed a bifunctor.
Finally, complete the category $\tilde\mathcal{P}$ with respect to
subobjects. This is our C$^*$-tensor category~$\mathcal{P}$, possibly with
nonsimple unit. The unitary tensor functor $\Pi\colon\mathcal{C}\to\mathcal{P}$ is
defined in the obvious way: it is the strict tensor functor which is
the identity map on objects and $\Pi(T)=(\iota_X\otimes T)_X$ on
morphisms. We will often omit $\Pi$ and simply consider $\mathcal{C}$ as a
C$^*$-tensor subcategory of $\mathcal{P}$.
\begin{definition} The pair $(\mathcal{P},\Pi)$ is called the {\em Poisson
boundary} of $(\mathcal{C},\mu)$. We say that the Poisson boundary is
trivial if $\Pi\colon\mathcal{C}\to\mathcal{P}$ is an equivalence of categories, or
in other words, for all objects~$U$ and~$V$ in $\mathcal{C}$ the only
bounded $P_\mu$-harmonic natural transformations $\iota\otimes
U\to\iota\otimes V$ are the transformations of the form
$\eta=(\iota_X\otimes T)_X$ for $T\in\mathcal{C}(U,V)$.
\end{definition}
The algebra $\mathcal{P}(\mathds{1})$ is determined by the random walk on $\Irr(\mathcal{C})$
with transition probabilities
$$
p_\mu(s,t)=\sum_r\mu(r)m^t_{rs}\frac{d(t)}{d(r) d(s)},
$$
where $d(s)=d(U_s)$ and $m^t_{rs}=\dim\mathcal{C}(U_t,U_r\otimes
U_s)$. Namely, if we identify $\hat\mathcal{C}(\mathds{1})$ with
$$
\ell^\infty\text{-}\bigoplus_s\mathcal{C}(U_s)=\ell^\infty(\Irr(\mathcal{C})),
$$
then the operator $P_\mu$ on $\hat\mathcal{C}(\mathds{1})$ is the Markov operator
defined by $p_\mu$, so $(P_\mu f)(s)=\sum_tp_\mu(s,t)f(t)$. Therefore
$\mathcal{P}(\mathds{1})$ is the algebra of bounded measurable functions on the
Poisson boundary, in the usual probabilistic sense, of the random walk
on $\Irr(\mathcal{C})$ with transition probabilities~$p_\mu(s,t)$. We say
that~$\mu$ is \emph{ergodic}, if this boundary is trivial, that is,
the tensor unit of~$\mathcal{P}$ is simple.
We say that $\mu$ is \emph{symmetric} if $\mu(s)=\mu(\bar s)$ for all
$s$, and that $\mu$ is \emph{generating} if every simple object
appears in the decomposition of $U_{s_1}\otimes\dots \otimes U_{s_n}$
for some $s_1,\dots,s_n\in\supp\mu$ and $n\ge1$. Equivalently, $\mu$
is generating if $\cup_{n\ge1}\supp\mu^{*n}=\Irr(\mathcal{C})$, where the
convolution of probability measures on $\Irr(\mathcal{C})$ is defined by
$$
(\nu*\mu)(t)=\sum_{s,r}\nu(s)\mu(r)m^t_{sr}\frac{d(t)}{d(s) d(r)}.
$$
We will write $\mu^n$ instead of $\mu^{*n}$. The definition of the
convolution is motivated by the identity $P_\mu P_\nu=P_{\nu*\mu}$.
We remark that a symmetric ergodic measure $\mu$, or even an ergodic
measure with symmetric support, is automatically generating. Indeed,
the symmetry assumption implies that we have a well-defined
equivalence relation on $\Irr(\mathcal{C})$ such that $s\sim t$ if and only if
$t$ can be reached from $s$ with nonzero probability in a finite
nonzero number of steps. Then any bounded function on $\Irr(\mathcal{C})$ that
is constant on equivalence classes is $P_\mu$-harmonic. Hence $\mu$ is
generating by the ergodicity assumption.
\smallskip
Let us say that $\mathcal{C}$ is \emph{weakly amenable} if the fusion algebra
$(\mathbb{Z}[\Irr(\mathcal{C})],d)$ is weakly amenable in the sense of Hiai and
Izumi~\cite{MR1644299}, that is, there exists a left invariant mean on
$\ell^\infty(\Irr(\mathcal{C}))$. By definition this is a state $m$ such that
$m(P_s(f))=m(f)$ for all $f\in\ell^\infty(\Irr(\mathcal{C}))$ and
$s\in\Irr(\mathcal{C})$. Of course, it is also possible to define right
invariant means, and by \cite{MR1644299}*{Proposition~4.2} if there
exists a left or right invariant mean, then there exists a
bi-invariant mean. By the same proposition amenability implies weak
amenability, as the term suggests. But as opposed to the group case,
in general, the converse is not true. Using this terminology let us
record the following known result.
\begin{proposition} \label{pweakamen} An ergodic probability measure
on $\Irr(\mathcal{C})$ exists if and only if $\mathcal{C}$ is weakly
amenable. Furthermore, if an ergodic measure exists, then it can be
chosen to be symmetric and with support equal to the entire space
$\Irr(\mathcal{C})$.
\end{proposition}
\begin{proof} If $\mu$ is an ergodic measure, then any weak$^*$ limit point of
the sequence $n^{-1}\sum^{n-1}_{k=0}\mu^k$ defines a right invariant
mean. For random walks on groups this implication was observed by
Furstenberg. The other direction is proved
in~\cite{MR1749868}*{Theorem~2.5}. It is an analogue of a result of
Kaimanovich--Vershik and Rosenblatt. \end{proof}
It should be remarked that if the fusion algebra of $\mathcal{C}$ is weakly
amenable and finitely generated, in general it is not possible to find
a finitely supported ergodic
measure~\cite{MR704539}*{Proposition~6.1}.
\smallskip
To finish the section, let us show that, not surprisingly, categorical
Poisson boundaries are of interest only for infinite categories.
\begin{proposition}\label{pfb} Assume $\mathcal{C}$ is finite, meaning that
$\Irr(\mathcal{C})$ is finite, and $\mu$ is generating. Then the Poisson
boundary of $(\mathcal{C},\mu)$ is trivial.
\end{proposition}
\begin{proof} The proof is similar to the proof of triviality of the Poisson
boundary of a random walk on a finite set based on the maximum
principle. Fix an object $U$ in $\mathcal{C}$ and assume that
$\eta\in\hat{\mathcal{C}}(U)$ is positive and $P_\mu$-harmonic. We claim that
if $\eta\ne0$ then there exists a positive nonzero morphism
$T\in\End{\mathcal{C}}{U}$ such that $\eta\ge T$. Assuming that the claim is
true, we can then choose a maximal $T$ with this property. Applying
again the claim to the element $\eta-T$, we conclude that $\eta=T$ by
maximality.
In order to prove the claim observe that $\eta_e\in\End{\mathcal{C}}{U}$ is
nonzero. Indeed, by assumption there exists~$s$ such that
$\eta_s\ne0$. Since the categorical traces are faithful, and therefore
partial categorical traces are faithful completely positive maps, it
follows that $P_s(\eta)_e\ne0$. Since $s\in\supp\mu^n$ for some
$n\ge1$, we conclude that $\eta_e=P_{\mu^n}(\eta)_e\ne0$.
Denote the positive nonzero element $\eta_e\in\End{\mathcal{C}}{U}$ by
$S$. Fix $s\in\Irr(\mathcal{C})$. Let $(R_s,\bar R_s)$ be a standard solution
of the conjugate equations for $U_s$, and $p\in \End{\mathcal{C}}{\bar
U_s\otimes U_s}$ be the projection defined by $p=d(s)^{-1}R_sR_s^*$. By
naturality of $\eta$ we then have $\eta_{\bar U_s\otimes U_s}\ge
p\otimes S$, whence
$$
P_{\bar s}(\eta)_s\ge(\tr_{\bar s}\otimes\iota)(p)\otimes
S=d(s)^{-2}(\iota\otimes S).
$$
Using the generating property of $\mu$ and finiteness of $\Irr(\mathcal{C})$,
we conclude that there exists a number $\lambda>0$ such that
$\eta_s\ge \iota\otimes\lambda S$ for all $s$. This proves the
claim. \end{proof}
\bigskip
\section{Realizations of the Poisson boundary}
As in the previous section, we fix a strict rigid C$^*$-tensor
category $\mathcal{C}$ and a probability measure~$\mu$ on~$\Irr(\mathcal{C})$. In
Sections~\ref{sec:longo-roberts-appr} and~\ref{sHY} we will in
addition assume that $\mu$ is generating. Let $\Pi\colon\mathcal{C}\to\mathcal{P}$ be
the Poisson boundary of $(\mathcal{C},\mu)$. Our goal is to give several
descriptions of the algebras~$\mathcal{P}(U)$ of harmonic elements.
\subsection{Time shift on the categorical path space}
\label{sec:poiss-bound-from}
Fix an object $U$. Denote by $M^{(0)}_U$ the von Neumann algebra
$\hat{\mathcal{C}}(U)\cong\ell^\infty{\text-}\oplus_s\End{\mathcal{C}}{U_s\otimes
U}$. More generally, for every $n\ge0$ consider the von Neumann
algebra
$$
M^{(n)}_U=\Endb(\iota_{\mathcal{C}^{n+1}}\otimes U),
$$
so $M^{(n)}_U$
consists of bounded collections
$\eta=(\eta_{X_{n},\dots,X_0})_{X_n,\dots,X_0}$ of natural in
$X_n,\dots,X_0$ endomorphisms of $X_n\otimes\dots \otimes X_0\otimes
U$. We consider $M^{(n)}_U$ as a subalgebra of~$M^{(n+1)}_U$ using the
embedding
$$
(\eta_{X_{n},\dots,X_0})_{X_n,\dots,X_0}\mapsto(\iota_{X_{n+1}}\otimes
\eta_{X_n,X_{n-1},\dots,X_0})_{X_{n+1},\dots,X_0}.
$$
Define a conditional
expectation $E_{n+1,n}\colon M^{(n+1)}_U\to M^{(n)}_U$ by
$$
E_{n+1,n}(\eta)_{X_n,\dots,X_0} = \sum_s \mu(s)(\tr_s\otimes\iota)
(\eta_{U_s,X_n,\dots,X_0}).
$$
Taking compositions of such conditional expectations we get normal
conditional expectations $$E_{n,0}\colon M^{(n)}_U\to M^{(0)}_U.$$
These conditional expectations are not faithful for $n\ge1$ unless the
support of $\mu$ is the entire space $\Irr(\mathcal{C})$. The support of
$E_{n,0}$ is a central projection, and we denote by $\mathcal{M}^{(n)}_U$ the
reduction of $M^{(n)}_U$ by this projection. More concretely, we have
a canonical isomorphism
\begin{equation}\label{eq:M-n-U-defn}
\mathcal{M}^{(n)}_U \cong \ell^\infty\text{-}\bigoplus_{\substack{s_n,\dots,s_1\in\supp\mu\\ s_0\in\Irr(\mathcal{C})}}
\End{\mathcal{C}}{U_{s_n}\otimes\dots\otimes U_{s_0}\otimes U}.
\end{equation} The conditional expectations $E_{n,0}$ define normal
faithful conditional expectations
$\mathcal{E}_{n,0}\colon\mathcal{M}^{(n)}_U\to\mathcal{M}^{(0)}_U=M^{(0)}_U$, and similarly
$E_{n+1,n}$ define conditional expectations $\mathcal{E}_{n+1,n}$. Denote by
$\mathcal{M}_U$ the von Neumann algebra obtained as the inductive
limit of the algebras $\mathcal{M}^{(n)}_U$ with respect to $\mathcal{E}_{n,0}$. In other
words, take any faithful normal state $\phi^{(0)}_U$ on $\mathcal{M}^{(0)}_U$. By
composing it with the conditional expectation $\mathcal{E}_{n,0}$ we get a
state $\phi^{(n)}_U$ on $\mathcal{M}^{(n)}_U$. Together these states define a
state on $\cup_n\mathcal{M}^{(n)}_U$. Finally, complete $\cup_n\mathcal{M}^{(n)}_U$ to a
von Neumann algebra in the GNS-representation corresponding to this
state. Denote the corresponding normal state on $\mathcal{M}_U$
by~$\phi_U$.
Note that if we start with a trace on
$\mathcal{M}^{(0)}_U$ which is a convex combination of the traces
$\tr_{U_s\otimes U}$, then the corresponding state $\phi_U$
on $\mathcal{M}_U$ is tracial. Since it is faithful on $\mathcal{M}^{(n)}_U$
for every~$n$, it is faithful on~$\mathcal{M}_U$. This shows that
$\mathcal{M}_U$ is a finite von Neumann algebra. Furthermore, the
$\phi_U$-preserving normal faithful conditional expectation
$\mathcal{E}_n\colon\mathcal{M}_U\to \mathcal{M}^{(n)}_U$ coincides with $\mathcal{E}_{n+1,n}$
on $\mathcal{M}^{(n+1)}_U$. It follows that on the dense algebra
$\cup_m\mathcal{M}^{(m)}_U$ the conditional expectation $\mathcal{E}_n$ is the limit, in
the pointwise $s^*$-topology, of $\mathcal{E}_{n+1,n}\mathcal{E}_{n+2,n+1}\dots
\mathcal{E}_{m+1,m}$ as $m\to\infty$. Hence $\mathcal{E}_n$ is independent of the choice
of a faithful normal trace $\phi^{(0)}_U$ as above.
\smallskip
Define a unital endomorphism $\theta_U$ of
$\cup_nM^{(n)}_U$ such that $\theta_U(M^{(n)}_U)\subset M^{(n+1)}_U$
by
$$
\theta_U(\eta)_{X_{n+1},\dots,X_0}=\eta_{X_{n+1},\dots,X_2,X_1\otimes
X_0}.
$$
Considering $\mathcal{M}^{(k)}_U$ as a quotient of $M^{(k)}_U$ we get a unital
endomorphism of $\cup_n\mathcal{M}^{(n)}_U$.
\begin{lemma} \label{lshift}
The endomorphism $\theta_U$ of $\cup_n\mathcal{M}^{(n)}_U$
extends to a normal faithful endomorphism of~$\mathcal{M}_U$, which
we continue to denote by $\theta_U$.
\end{lemma}
\begin{proof} Consider the normal semifinite faithful (n.s.f.)~trace
$\psi^{(0)}_U=\sum_sd(s)^2\tr_{U_s\otimes U}$ on
$$\mathcal{M}^{(0)}_U\cong\ell^\infty\text{-}\bigoplus_s\End{\mathcal{C}}{U_s\otimes U}$$ and put
$\psi_U=\psi^{(0)}_U\mathcal{E}_0$. Then $\psi_U$ is an n.s.f.~trace. In order
to prove the lemma it suffices to show that the restriction of
$\psi_U$ to $\cup_n\mathcal{M}^{(n)}_U$ is $\theta_U$-invariant. Indeed, if the
invariance holds, then we can define an isometry $U$ on
$L^2(\mathcal{M}_U,\psi_U)$ by
$U\Lambda_{\psi_U}(x)=\Lambda_{\psi_U}(\theta_U(x))$ for
$x\in\cup_n\mathcal{M}^{(n)}_U$ such that $\psi_U(x^*x)<\infty$. Let $H\subset
L^2(\mathcal{M}_U,\psi_U)$ be the image of~$U$ and $\mathcal{M}$ be the von Neumann
algebra generated by the image of $\theta_U$. Then $H$ is
$\mathcal{M}$-invariant. We can choose $0\le e_i\le1$ such that
$\theta_U(e_i)\to1$ strongly and $\psi_U(e_i)<\infty$. Now, if
$x\in\mathcal{M}_+$ is such that $x|_H=0$, then
$\psi_U(\theta_U(e_i)x\theta_U(e_i))=0$, and by lower semicontinuity
we get $\psi_U(x)=0$, so $x=0$. Therefore we can define $\theta_U$ as
the composition of the map $\mathcal{M}_U\to B(H)$, $x\mapsto UxU^*$, with the
inverse of the map $\mathcal{M}\to \mathcal{M}|_H$.
It remains to check the invariance. By definition we have
$\mathcal{E}_{n+2,n+1}\theta_U=\theta_U\mathcal{E}_{n+1,n}$ on $\mathcal{M}^{(n+1)}_U$ for all
$n\ge0$. This implies that $\mathcal{E}_{n+1}\theta_U=\theta_U\mathcal{E}_n$ on
$\cup_k\mathcal{M}^{(k)}_U$. It follows that for any $x\in \cup_n\mathcal{M}^{(n)}_U$ we
have
$$
\psi_U\theta_U(x)=\psi_U\mathcal{E}_1\theta_U(x)=\psi_U\theta_U\mathcal{E}_0(x)
=\psi_U\mathcal{E}_0\theta_U\mathcal{E}_0(x).
$$
This implies that it suffices to show that $\psi_U\mathcal{E}_0\theta_U=\psi_U$
on~$\mathcal{M}^{(0)}_U$. Since $\tr_{U_s\otimes U}=\tr_s(\iota\otimes\tr_U)$,
it is enough to consider the case $U=\mathds{1}$. Note also that
$\mathcal{E}_0\theta_U=P_\mu$ on~$\mathcal{M}^{(0)}_U$. Thus we have to check that
$\psi_\mathds{1} P_\mu=\psi_\mathds{1}$ on
$\mathcal{M}^{(0)}_\mathds{1}\cong\ell^\infty(\Irr(\mathcal{C}))$. This is equivalent to the
easily verifiable identity $\mu*m=m$, where
$m=\sum_sd(s)^2\delta_s$. \end{proof}
We call the endomorphism $\theta_U$ of $\mathcal{M}_U$ the {\em time
shift}. Now, take $\eta\in\mathcal{M}^{(0)}_U$. Then for every $n\ge0$ we can
define an element $\eta^{(n)}\in M^{(n)}_U$ by
$$
\eta^{(n)}_{X_n,\dots,X_0}=\eta_{X_n\otimes\dots\otimes X_0}.
$$
Consider the image of $\eta^{(n)}$ in $\mathcal{M}^{(n)}_U$ and denote it again
by $\eta^{(n)}$, since this is the only element we are interested
in. Then $\eta$ is $P_\mu$-harmonic if and only if
$\mathcal{E}_{1,0}(\eta^{(1)})=\eta$, and in this case
$\mathcal{E}_{n+1,n}(\eta^{(n+1)})=\eta^{(n)}$ for all $n$. Therefore if $\eta$
is $P_\mu$-harmonic, then the sequence $\{\eta^{(n)}\}_n$ is a
martingale. Denote by $\eta^{(\infty)}\in\mathcal{M}_U$ its $s^*$-limit.
\begin{proposition} \label{prop:lr-emb} The map
$\eta\mapsto\eta^{(\infty)}$ is an isomorphism between the von Neumann
algebra $\mathcal{P}(U)$ of $P_\mu$-harmonic bounded natural transformations
$\iota\otimes U\to\iota\otimes U$ and the fixed point algebra
$\mathcal{M}^{\theta_U}_U$. The inverse map is given by $x\mapsto \mathcal{E}_0(x)$.
\end{proposition}
\begin{proof} By definition we have $\eta^{(n)}=\theta^n_U(\eta)$. It follows
that if $\eta$ is $P_\mu$-harmonic, so that
$\eta^{(n)}\to\eta^{(\infty)}$, then the element $\eta^{(\infty)}$ is
$\theta_U$-invariant. We also clearly have
$\mathcal{E}_0(\eta^{(\infty)})=\eta$.
Conversely, take $x\in\mathcal{M}_U^{\theta_U}$. The proof of
Lemma~\ref{lshift} implies that $\mathcal{E}_{n+1}\theta_U=\theta_U\mathcal{E}_n$. Hence
the martingale $\{x_n=\mathcal{E}_n(x)\}_n$ has the property
$x_{n+1}=\theta_U(x_n)$. As $\mathcal{E}_0\theta_U=P_\mu$ on~$\mathcal{M}^{(0)}_U$, we
conclude that $x_0$ is $P_\mu$-harmonic and $x_0^{(\infty)}=x$.
We have thus proved that the maps in the formulation are inverse to
each other. Since they are unital completely positive, they must be
isomorphisms. \end{proof}
The bijection between $\mathcal{P}(U)$ and $\mathcal{M}^{\theta_U}_U$ could be used to
give an alternative proof of Proposition~\ref{pproduct}. Namely, we
could define a product $\cdot$ on harmonic elements by
$\nu\cdot\eta=\mathcal{E}_0(\nu^{(\infty)}\eta^{(\infty)})$. Since
$\nu^{(\infty)}\eta^{(\infty)}$ is the $s^*$-limit of the elements
$\nu^{(n)}\eta^{(n)}=(\nu\eta)^{(n)}$, and
$\mathcal{E}_0((\nu\eta)^{(n)})=P^n_\mu(\nu\eta)$, it follows that
$P^n_\mu(\nu\eta)\to \nu\cdot\eta$ in the $s^*$-topology, which is
equivalent to saying that $P^n_\mu(\nu\eta)_X\to (\nu\cdot\eta)_X$ for
every $X$.
\subsection{Relative commutants: Izumi--Longo--Roberts approach}
\label{sec:longo-roberts-appr}
We will now modify the construction of the algebras $\mathcal{M}_U$ to get
algebras $\mathcal{N}_U$ and an identification of $\mathcal{P}(U)$ with
$\mathcal{N}_\mathds{1}'\cap\mathcal{N}_U$. Conceptually, instead of considering all paths
of the random walk defined by $\mu$, we consider only paths starting
at the unit object. The time shift is no longer defined on this space,
but by considering a larger space we can still get a description of
$\mathcal{P}(U)$ in simple von Neumann algebraic terms. For this to work we
have to assume that $\mu$ is generating, so that we can reach any
simple object from the unit.
This identification of harmonic elements is closely related to Izumi's
description of Poisson boundaries of discrete quantum
groups~\cite{MR1916370}. A similar construction was also used by Longo
and Roberts using sector theory~\cite{MR1444286}. More precisely, they
worked with a somewhat limited form of $\mu$ and what we obtain is a
possibly infinite von Neumann algebra for what corresponds to the
finite gauge-invariant von Neumann subalgebra in their work.
\smallskip
We first put $V = \oplus_{s \in \supp \mu} U_s$. In the case $\supp
\mu$ is infinite, this should be understood only as a suggestive
notation which does not make sense inside $\mathcal{C}$. Given an object $U$,
by $\mathcal{C}(V^{\otimes n}\otimes U)$ we understand the space
$$
\bigoplus_{s_*, s'_* \in \supp \mu^n} \mathcal{C}(U_{s_n} \otimes \cdots
\otimes U_{s_1} \otimes U, U_{s'_n} \otimes \cdots\otimes U_{s'_1}
\otimes U)
$$
endowed with the obvious $*$-algebra structure. Similarly to
Section~\ref{sec:poiss-bound-from} we have completely positive maps
$$
\mathcal{E}_{n+1,n}=\sum_s \mu(s) (\tr_s \otimes \iota)\colon \mathcal{C}(V^{\otimes
(n+1)}\otimes U)\to \mathcal{C}(V^{\otimes n}\otimes U),
$$
and taking the composition of these maps we get maps
$$
\mathcal{E}_{n,0}\colon \mathcal{C}(V^{\otimes n}\otimes U)\to \mathcal{C}(U).
$$
Then $\omega_U^{(n)} = \tr_U \mathcal{E}_{n,0}$ is a state on $\mathcal{C}(V^{\otimes
n}\otimes U)$. We denote by $\mathcal{N}^{(n)}_U$ the von Neumann algebra
generated by $\mathcal{C}(V^{\otimes n}\otimes U)$ in the GNS-representation
defined by this state. The elements of $\mathcal{N}_U^{(n)}$ are represented
by certain bounded families in the direct product of the morphism
sets $$\mathcal{C}(U_{s_n} \otimes \cdots \otimes U_{s_1} \otimes U, U_{s'_n}
\otimes \cdots \otimes U_{s'_1} \otimes U).$$ Since the positive
elements of $\mathcal{N}_U^{(n)}$ have positive diagonal entries, the state
$\omega_U^{(n)}$ is faithful on $\mathcal{N}_U^{(n)}$.
There is a natural diagonal embedding $\mathcal{N}_U^{(n)} \rightarrow
\mathcal{N}_U^{(n+1)}$ defined by $T \mapsto \iota_V \otimes T$. The map
$\mathcal{E}_{n+1,n}$ extends then to a normal conditional expectation
$\mathcal{N}_U^{(n+1)} \rightarrow \mathcal{N}_U^{(n)}$ such that
$\omega^{(n)}_U\mathcal{E}_{n+1,n}=\omega^{(n+1)}_U$. This way we obtain an
inductive system $(\mathcal{N}_U^{(n)}, \omega_U^{(n)})_n$ of von Neumann
algebras, and we let $(\mathcal{N}_U, \omega_U)$ be the von Neumann algebra
and the faithful state obtained as the limit. As in
Section~\ref{sec:poiss-bound-from}, composing the conditional
expectations $\mathcal{E}_{n+1,n}$ and passing to the limit we get
$\omega_U$-preserving conditional expectations
$\mathcal{E}_n\colon\mathcal{N}_U\to\mathcal{N}^{(n)}_U$.
\smallskip
When $U = \mathds{1}$, we simply write $\mathcal{N}^{(n)}$ and $\mathcal{N}$ instead of
$\mathcal{N}^{(n)}_\mathds{1}$ and $\mathcal{N}_\mathds{1}$. If $U'$ and $U$ are objects in~$\mathcal{C}$,
then the map $x\mapsto x\otimes\iota_U$ defines an embedding
$\mathcal{N}_{U'}\hookrightarrow\mathcal{N}_{U'\otimes U}$. In particular, the
algebra $\mathcal{N}$ is contained in any of $\mathcal{N}_U$.
\smallskip
When $\eta$ is a natural transformation in $\hat\mathcal{C}(U)$,
the morphism $$\eta_{V^{\otimes n}} = \oplus_{s_*} \eta_{U_{s_n}
\otimes \cdots \otimes U_{s_1}}$$ defines an element in the diagonal
part of $\mathcal{N}_U^{(n)}$, which we denote by $\eta^{[n]}$. Note that
the direct summand $s_0 = e$ of~\eqref{eq:M-n-U-defn} can be
identified with the diagonal part of $\mathcal{N}_U^{(n)}$, and $\eta^{[n]}$
simply becomes the component of $\eta^{(n)}$ in this summand. If
$\eta$ is $P_\mu$-harmonic, the sequence $\{\eta^{[n]}\}_n$ forms a
martingale and defines an element $\eta^{[\infty]} \in
\mathcal{N}_U^{(\infty)}$.
\begin{proposition}
\label{prop:p-bdry-as-rel-comm-LR0} For every object $U$ in $\mathcal{C}$, the
map $\eta \mapsto\eta^{[\infty]}$ defines an isomorphism of von
Neumann algebras $\End{\mathcal{P}}{U}\cong \mathcal{N}'\cap\mathcal{N}_{U}$.
\end{proposition}
\begin{proof} If $\eta$ is a harmonic element in $\hat\mathcal{C}(U)$, the
naturality implies that the elements $\eta_{V^{\otimes m}}$ commute
with the image of $\End{\mathcal{C}}{V^{\otimes n}}$ for $m \ge n$. Thus,
$\eta^{[\infty]} = \lim_m \eta_{V^{\otimes m}}$ is in the relative
commutant. Since~$\mu$ is generating, it is also clear that the map
$\eta \mapsto \eta^{[\infty]}$ is injective.
To construct the inverse map, take an element
$x\in\mathcal{N}'\cap\mathcal{N}_{U}$. Then $x_n = \mathcal{E}_n(x)$ is an element of
$(\mathcal{N}^{(n)})' \cap \mathcal{N}_{U}^{(n)}$. Hence, for every $n\ge1$ and
$s\in\supp\mu^n$, there is a morphism $x_{n, s} \in \End{\mathcal{C}}{U_s
\otimes U}$ such that $x_n$ is the direct sum of the $x_{n, s}$ (with
multiplicities). It follows that we can choose $\eta(n)\in \hat\mathcal{C}(U)$
such that $\|\eta(n)\|\le\|x\|$ and $x_n=\eta(n)^{[n]}$. The elements
$\eta(n)$ are not uniquely determined, only their components
corresponding to $s\in\supp\mu^n$ are. The identity $\mathcal{E}_{n+1,
n}(x_{n+1}) = x_n$ translates into $P_\mu(\eta(n+1))_s=\eta(n)_s$ for
$s\in\supp\mu^n$.
We now define an element $\eta\in\Endb(U)$ by letting
$$
\eta_s=\eta(n)_s\ \ \text{if}\ \ s\in\supp\mu^n\ \ \text{for some}\ \
n\ge1.
$$
In order to see that this definition in unambiguous, assume
$s\in(\supp\mu^n)\cap(\supp\mu^{n+k})$ for some~$n$ and~$k$. Then by
the $0$-$2$ law, see~\cite{MR2034922}*{Proposition~2.12}, we have
$\|P^m_\mu-P^{m+k}_\mu\|\to0$ as $m\to\infty$. Since the sequence
$\{\eta(m)\}_m$ is bounded and we have
$\eta(n)_s=P^{m+k}_\mu(\eta(n+m+k))_s$ and
$\eta(n+k)_s=P^{m}_\mu(\eta(n+m+k))_s$, letting $m\to\infty$ we
conclude that $\eta(n)_s=\eta(n+k)_s$. Hence $\eta$ is well-defined,
$P_\mu$-harmonic, and $x_n=\eta^{[n]}$. Therefore $x=\eta^{[\infty]}$.
The linear isomorphism $\mathcal{P}(U)\to\mathcal{N}'\cap\mathcal{N}_U$ and its inverse that
we have constructed, are unital and completely positive, hence they
are isomorphisms of von Neumann algebras. \end{proof}
As in the case of Proposition~\ref{prop:lr-emb}, the linear
isomorphism $\mathcal{P}(U)\cong\mathcal{N}'\cap\mathcal{N}_U$ could be used to give an
alternative proof of Proposition~\ref{pproduct}, at least for
generating measures.
\smallskip
Applying Proposition~\ref{prop:p-bdry-as-rel-comm-LR0} to $U=\mathds{1}$ we
get the following.
\begin{corollary}\label{cor:ergod-factor} The von Neumann algebra
$\mathcal{N}$ is a factor if and only if $\mu$ is ergodic.
\end{corollary}
Under a mildly stronger assumption on the measure we can prove a
better result than Proposition~\ref{prop:p-bdry-as-rel-comm-LR0},
which will be important later.
\begin{proposition}
\label{prop:p-bdry-as-rel-comm-LR} Assume that for any
$s,t\in\Irr(\mathcal{C})$ there exists $n\ge0$ such
that $$\supp(\mu^n*\delta_s)\cap\supp(\mu^n*\delta_t)\ne\emptyset.$$
Then for any objects $U$ and $U'$ in $\mathcal{C}$, the map $\eta \mapsto
(\iota_{U'} \otimes \eta)^{[\infty]}$ defines an isomorphism of von
Neumann algebras $\End{\mathcal{P}}{U}\cong \mathcal{N}_{U'}'\cap\mathcal{N}_{U' \otimes
U}$.
\end{proposition}
\begin{proof} That we get a map $\End{\mathcal{P}}{U}\to \mathcal{N}_{U'}'\cap\mathcal{N}_{U' \otimes
U}$ does not require any assumptions on $\mu$ and is easy to see: if
$\eta$ is a harmonic element in $\hat\mathcal{C}(U)$, the
naturality implies that the elements $\eta_{V^{\otimes m} \otimes U'}$
commute with $\End{\mathcal{C}}{V^{\otimes n} \otimes U'}$ for $m \ge n$, and
hence $(\iota_{U'} \otimes \eta)^{[\infty]} = \lim_m \eta_{V^{\otimes
m} \otimes U'}$ lies in $ \mathcal{N}_{U'}'\cap\mathcal{N}_{U' \otimes U}$.
To construct the inverse map assume first $U'=U_t$ for some $t$. Take
$x\in \mathcal{N}_{U'}'\cap\mathcal{N}_{U' \otimes U}$. Similarly to the proof of
Proposition~\ref{prop:p-bdry-as-rel-comm-LR0} we can find elements
$\eta(n)\in\hat{\mathcal{C}}(U)$ such that $\|\eta(n)\|\le\|x\|$ and
$\mathcal{E}_n(x)=(\iota_{U'}\otimes\eta(n))^{[n]}$. The identity $\mathcal{E}_{n+1,
n}(x_{n+1}) = x_n$ means now that $P_\mu(\eta(n+1))_s=\eta(n)_s$ for
$s\in\supp(\mu^n*\delta_t)$. We want to define an element
$\eta\in\hat\mathcal{C}(U)$ by
$$
\eta_s=\eta(n)_s\ \ \text{if}\ \ s\in\supp(\mu^n*\delta_t)\ \
\text{for some}\ \ n\ge1.
$$
As in the proof of Proposition~\ref{prop:p-bdry-as-rel-comm-LR0}, in
order to see that $\eta$ is well-defined, it suffices to show that if
$s\in\supp(\mu^n*\delta_t)\cap\supp(\mu^{n+k}*\delta_t)$ for some~$n$
and~$k$, then $\|P^m_\mu-P^{m+k}_\mu\|\to0$ as $m\to\infty$. Since
$\mu$ is assumed to be generating, there exists $l$ such that
$t\in\supp\mu^l$. But then
$$
s\in (\supp\mu^{n+l})\cap(\supp\mu^{n+l+k}),
$$
so the convergence $\|P^m_\mu-P^{m+k}_\mu\|\to0$ indeed holds by the
$0$-$2$ law. This finishes the proof of the proposition for $U'=U_t$,
and we see that no assumption in addition to the generating property
of $\mu$ is needed in this case.
\smallskip
Consider now an arbitrary $U'$. Decompose $U'$ into a direct sum of
simple objects:
$$
U\cong U_{s_1}\oplus\dots\oplus U_{s_n}.
$$
Denote by $p_i\in\mathcal{C}(U')$ the corresponding projections. Then the
inclusion $p_i\mathcal{N}_{U'}p_i\subset p_i\mathcal{N}_{U'\otimes U}p_i$ can be
identified with $\mathcal{N}_{U_{s_i}}\subset \mathcal{N}_{U_{s_i}\otimes U}$.
Take $x\in \mathcal{N}_{U'}'\cap\mathcal{N}_{U'\otimes U}$. Then $x$ commutes with
$p_i$. Since the element $xp_i$ lies in
$\mathcal{N}_{U_{s_i}}'\cap\mathcal{N}_{U_{s_i}\otimes U}$, it is defined by a
$P_\mu$-harmonic element $\eta(i)\in\hat{\mathcal{C}}(U)$. In terms
of these elements the condition that $\mathcal{E}_n(x)$ commutes with
$\mathcal{C}(V^{\otimes n}\otimes U')$ means that $\eta(i)_s=\eta(j)_s$
whenever $s\in\supp(\mu^n*\delta_{s_i})\cap\supp(\mu^n*\delta_{s_j})$,
while to finish the proof we need the equality $\eta(i)=\eta(j)$.
Fix $s\in\Irr(\mathcal{C})$ and indices $i$ and $j$. By assumption there
exists $t\in\supp(\mu^n*\delta_{s_i})\cap\supp(\mu^n*\delta_{s_j})$
for some $n$. Since $\mu$ is generating, there exists $m$ such that
$s\in\supp(\mu^m*\delta_t)$. Then
$$
s\in \supp(\mu^{m+n}*\delta_{s_i})\cap\supp(\mu^{m+n}*\delta_{s_j}),
$$
and therefore $\eta(i)_s=\eta(j)_s$. \end{proof}
Note that the proof shows that the additional assumption on the
measure is not only sufficient but is also necessary for the result to
be true. Even for symmetric ergodic measures this condition does not
always hold: take the random walk on $\mathbb{Z}$ defined by the measure
$\mu=2^{-1}(\delta_{-1}+\delta_1)$. At the same time this condition is
satisfied, for example, for any generating measure $\mu$ with
$\mu(e)>0$. Indeed, for such a measure we can find $n$ such that
$s\in\supp(\mu^n*\delta_t)$, and then $s\in\supp(\mu^n*\delta_s)\cap
\supp(\mu^n*\delta_t)$.
\smallskip
Applying the proposition to $U=\mathds{1}$ we get the following result.
\begin{corollary}\label{cor:engod-factor} Assume $\mu$ is ergodic and
satisfies the assumption of
Proposition~\ref{prop:p-bdry-as-rel-comm-LR}. Then $\mathcal{N}_U$ is a
factor for every object $U$ in $\mathcal{C}$.
\end{corollary}
\begin{remark}\label{rmultiplicity} It is sometimes convenient to
consider slightly more general constructions allowing
multiplicities. Namely, instead of $V=\oplus_{s\in\supp\mu}U_s$ we
could take $V=\oplus_{i\in I}U_{s_i}$, where $(s_i)_{i\in I}$ is any
finite or countable collection of elements running through
$\supp\mu$. For the state on $\mathcal{C}(V)$ we could take
$\mathcal{C}(U_{s_i},U_{s_j})\ni T \mapsto\delta_{ij}\lambda_i\tr_{s_i}(T)$,
where $\lambda_i>0$ are any numbers such that $\sum_{i\colon
s_i=s}\lambda_i=\mu(s)$ for all $s\in\supp\mu$. All the above results
would remain true, with essentially identical proofs.
\end{remark}
\subsection{Relative commutants: Hayashi--Yamagami
approach} \label{sHY}
We will now explain a modification of the Izumi--Longo--Roberts
construction due to Hayashi and Yamagami~\cite{MR1749868}. Its
advantage is that, at the expense of introducing an extra variable in
a II$_1$ factor, we can stay in the framework of finite von Neumann
algebras.
\smallskip
We continue to assume that $\mu$ is generating. We will use a slightly
different notation compared~\cite{MR1749868} to be more consistent
with the previous sections.
Let $\mathcal{R}$ be the hyperfinite II$_1$-factor and $\tau$ be the
unique normal tracial state on $\mathcal{R}$. Choose a partition of
unity by projections $(e_s)_{s \in \supp \mu}$ in $\mathcal{R}$ which satisfy
$$
\tau(e_s) = \frac{\mu(s)}{c d(s)},\ \ \text{where}\ \
c=\sum_{s\in\supp\mu}\frac{\mu(s)}{ d(s)}.
$$
When $(s_n, \ldots, s_1) \in (\supp \mu)^n$, we write $e_{s_*} =
e_{s_n} \otimes \cdots \otimes e_{s_1} \in \mathcal{R}^{\otimes
n}$. As in Section~\ref{sec:longo-roberts-appr}, put
$V=\oplus_{s\in\supp\mu}U_s$. Now, for a fixed object $U$ in $\mathcal{C}$,
instead of the algebra $\mathcal{C}(V^{\otimes n}\otimes U)$ used there,
consider the algebra
$$
\tilde\mathcal{C}(V^{\otimes n}\otimes U)=\bigoplus_{s_*, s'_* \in (\supp
\mu)^n} \mathcal{C}(U_{s_n} \otimes \cdots \otimes U_{s_1} \otimes U, U_{s'_n}
\otimes \cdots \otimes U_{s'_1} \otimes U) \otimes e_{s'_*}
\mathcal{R}^{\otimes n} e_{s_*}.
$$
It carries a tracial state $\tau^{(n)}_U$ defined by
$$
\tau^{(n)}_U(T\otimes x)=\delta_{s_*,s'_*}c^nd(s_1)\dots
d(s_n)\tr_{U_{s_n} \otimes\dots\otimes U_{s_1}\otimes
U}(T)\tau^{\otimes n}(x)
$$
for $T\otimes x\in \mathcal{C}(U_{s_n} \otimes \cdots \otimes U_{s_1} \otimes
U, U_{s'_n} \otimes \cdots \otimes U_{s'_1} \otimes U) \otimes
e_{s'_*} \mathcal{R}^{\otimes n} e_{s_*}$. Let $\mathcal{A}^{(n)}_U$ be the von
Neumann algebra generated by $\tilde\mathcal{C}(V^{\otimes n}\otimes U)$ in
the GNS-representation defined by $\tau^{(n)}_U$. These algebras form
an inductive system under the embeddings
$$
\mathcal{A}^{(n)}_U\hookrightarrow\mathcal{A}^{(n+1)}_U,\ \ T\otimes x\mapsto
\sum_{s\in\supp\mu}(\iota_s\otimes T)\otimes(e_s\otimes x).
$$
Passing to the limit we get a von Neumann algebra $\mathcal{A}_U$ equipped with
a faithful tracial state $\tau_U$. We write $\mathcal{A}$ for $\mathcal{A}_\mathds{1}$.
Given $\eta\in\hat{\mathcal{C}}(U)$, consider the elements
$$
\eta^{\{n\}}=\sum_{s_*\in(\supp \mu)^n} \eta_{U_{s_n} \otimes \cdots
\otimes U_{s_1}}\otimes e_{s_*}\in\mathcal{A}^{(n)}_U.
$$
If $\eta$ is $P_\mu$-harmonic, then the sequence $\{\eta^{\{n\}}\}_n$
forms a martingale with respect to the $\tau_U$-preserving conditional
expectations $\mathcal{E}_n\colon\mathcal{A}_U\to\mathcal{A}^{(n)}_U$. Denote its limit by
$\eta^{\{\infty\}}$. Then we get the following analogues of
Propositions~\ref{prop:p-bdry-as-rel-comm-LR0}
and~\ref{prop:p-bdry-as-rel-comm-LR}, with almost identical proofs,
which we omit.
\begin{proposition}\label{prop:PP-equals-HY} For every object $U$ in
$\mathcal{C}$, the map $\eta \mapsto\eta^{\{\infty\}}$ defines an isomorphism
of von Neumann algebras $\End{\mathcal{P}}{U}\cong \mathcal{A}'\cap\mathcal{A}_{U}$. If in
addition to the generating property the measure $\mu$ satisfies the
assumption of Proposition~\ref{prop:p-bdry-as-rel-comm-LR}, then also
the map $\eta \mapsto(\iota_{U'}\otimes\eta)^{\{\infty\}}$ defines an
isomorphism of von Neumann algebras $\End{\mathcal{P}}{U}\cong
\mathcal{A}_{U'}'\cap\mathcal{A}_{U'\otimes U}$ for any object $U'$.
\end{proposition}
The work of Hayashi and Yamagami contains much more than the
construction of the algebras $\mathcal{A}_U$ and, in fact, allows us to
describe, under mild additional assumptions on $\mu$, not only the
morphisms but the entire Poisson boundary $\Pi\colon\mathcal{C}\to\mathcal{P}$ in
terms of Hilbert bimodules over~$\mathcal{A}$.
For objects $X$ and $Y$ consider their direct sum $X\oplus Y$, and
denote by $p_X,p_Y\in\mathcal{C}(X\oplus Y)$ the corresponding projections. We
can consider $p_X$ and $p_Y$ as projections in $\mathcal{A}_{X\oplus Y}$, then
$p_X(\mathcal{A}_{X\oplus Y})p_X \cong \mathcal{A}_X$ and $p_Y(\mathcal{A}_{X\oplus Y})p_Y \cong
\mathcal{A}_Y$. Put
$$
\mathcal{A}_{X,Y}=p_Y(\mathcal{A}_{X\oplus Y})p_X.
$$
The $\mathcal{A}_Y$-$\mathcal{A}_X$-module $\mathcal{A}_{X,Y}$ can be described as an inductive
limit of completions of the spaces
$$
\tilde\mathcal{C}(V^{\otimes n}\otimes X,V^{\otimes n}\otimes
Y)=\bigoplus_{s_*, s'_* \in (\supp \mu)^n} \mathcal{C}(U_{s_n} \otimes \cdots
\otimes U_{s_1} \otimes X, U_{s'_n} \otimes \cdots \otimes U_{s'_1}
\otimes Y) \otimes e_{s'_*} \mathcal{R}^{\otimes n} e_{s_*}.
$$
Denote by $\mathcal{H}_X$ the Hilbert space completion of $\mathcal{A}_{\mathds{1},X}$ with
respect to the scalar product $$(x,y)=\tau_\mathds{1}(y^*x).$$ Then $\mathcal{H}_X$
is a Hilbert $\mathcal{A}_X$-$\mathcal{A}$-module (it is denoted by $X_\infty$
in~\cite{MR1749868}). Viewing $\mathcal{H}_X$ as a Hilbert bimodule over $\mathcal{A}$,
we get a unitary functor $F$ from $\mathcal{C}$ into the category $\mathrm{Hilb}_\mathcal{A}$
of Hilbert bimodules over~$\mathcal{A}$ such that $F(U)=\mathcal{H}_U$ on objects and
in the obvious way on morphisms in $\mathcal{C}$. We want to make~$F$ into a
tensor functor. By the computation on pp.~40--41 of~\cite{MR1749868}
the map
$$
\tilde\mathcal{C}(V^{\otimes n},V^{\otimes n}\otimes X)\otimes
\tilde\mathcal{C}(V^{\otimes n},V^{\otimes n}\otimes Y)\to\tilde
\mathcal{C}(V^{\otimes n},V^{\otimes n}\otimes X\otimes Y),
$$
$$
(S \otimes a)\otimes (T\otimes b)\mapsto (S\otimes\iota_Y)T \otimes
ab,
$$
defines an isometry
$$
F_2(X,Y)\colon \mathcal{H}_X\otimes_\mathcal{A}\mathcal{H}_Y\to\mathcal{H}_{X\otimes Y}.
$$
\begin{lemma} \label{lcondmeas} Assume that for every $s\in\Irr(\mathcal{C})$
we have
$$
(\mu^n*\delta_s)(\supp\mu^n)\to1\ \ \text{as}\ \ n\to\infty.
$$
Then the maps $F_2(X,Y)$ are unitary.
\end{lemma}
\begin{proof} It suffices to prove the lemma for simple objects. Assume $X=U_s$
for some $s$. For every $n\ge1$ and $s_*\in(\supp\mu)^n$, let
$p_{s_*}^{(n)}\in\mathcal{C}(U_{s_n}\otimes\dots\otimes U_{s_1}\otimes X)$ be
the projection onto the direct sum of the isotypic components
corresponding to $U_t$ for some $t\in\supp\mu^n$. Put
$$
p^{(n)}=\sum_{s_* \in (\supp \mu)^n}p^{(n)}_{s_*}\otimes e_{s_*}\in\mathcal{A}^{(n)}_{U_s}.
$$
Then $\tau_X(p^{(n)})=(\mu^n*\delta_s)(\supp\mu^n)$. Therefore by
assumption $p^{(n)}\to1$ in the $s^*$-topology. It follows that to
prove the lemma it suffices to show that if
$$
T\otimes x\in \mathcal{C}(U_{s_n} \otimes \cdots \otimes U_{s_1}, U_{s'_n}
\otimes \cdots \otimes U_{s'_1} \otimes X\otimes Y) \otimes e_{s'_*}
\mathcal{R}^{\otimes n} e_{s_*}
$$
is such that $p^{(n)}(T\otimes x)=T\otimes x$, then $T\otimes x$ is in
the image of $F_2(X,Y)$. The assumption on $T$ means that the simple
objects appearing in the decomposition of $U_{s'_n} \otimes \cdots
\otimes U_{s'_1} \otimes X$ appear also in the decomposition of
$U_{t_n}\otimes\dots\otimes U_{t_1}$ for $t_*\in(\supp\mu)^n$. This
implies that $T$ can be written as a finite direct sum of morphisms of
the form $(S\otimes\iota_Y)R$, with $R\in\mathcal{C}(U_{s_n} \otimes \cdots
\otimes U_{s_1},U_{t_n}\otimes\dots\otimes U_{t_1}\otimes Y)$ and
$S\in\mathcal{C}(U_{t_n} \otimes \cdots \otimes
U_{t_1},U_{s'_n}\otimes\dots\otimes U_{s'_1}\otimes Y)$. Since we also
have density of $e_{s'_*}\mathcal{R} e_{t_*}\mathcal{R} e_{s_*}$ in $e_{s'_*}\mathcal{R}
e_{s_*}$, this proves the lemma. \end{proof}
We remark that the assumption of the lemma is obviously satisfied if
$\supp\mu=\Irr(\mathcal{C})$. It is also satisfied if $\mu$ is ergodic and
$\mu(e)>0$, since then $\|\mu^n*\delta_s-\mu^n\|_1\to0$ by
\cite{MR1644299}*{Proposition~3.3}.
\smallskip
Once the maps $F_2(X,Y)$ are unitary, it is easy to see that $(F,F_2)$
is a unitary tensor functor $\mathcal{C}\to\mathrm{Hilb}_\mathcal{A}$.
\begin{proposition} \label{pHYrealization} Assume the measure $\mu$
satisfies the assumption of Lemma~\ref{lcondmeas}. Let $\mathcal{B}$ be the
full C$^*$-tensor subcategory of $\mathrm{Hilb}_\mathcal{A}$ generated by the image of
$F\colon\mathcal{C}\to\mathrm{Hilb}_\mathcal{A}$. Then the Poisson boundary
$\Pi\colon\mathcal{C}\to\mathcal{P}$ of $(\mathcal{C},\mu)$ is isomorphic to
$F\colon\mathcal{C}\to\mathcal{B}$.
\end{proposition}
\begin{proof} The functor $F$ extends to the full subcategory $\tilde\mathcal{P}$ of
$\mathcal{P}$ formed by the objects of $\mathcal{C}$ using the isomorphisms
$\mathcal{P}(U)\cong\mathcal{A}'\cap\mathcal{A}_U$. It follows immediately by definition that
this way we get a unitary tensor functor $E\colon\tilde\mathcal{P}\to\mathcal{B}$ if
we put $E_2(X,Y)=F_2(X,Y)$. We then extend this functor to a unitary
tensor functor $\mathcal{P}\to\mathcal{B}$, which we continue to denote by $E$. To
prove the proposition it remains to show that $E$ is fully
faithful. In other words, we have to show that the left action of
$\mathcal{A}_U$ on $\mathcal{H}_U$ defines an isomorphism
$\mathcal{A}'\cap\mathcal{A}_U\cong\Endd_{\mathcal{A}\mhyph\mathcal{A}}(\mathcal{H}_U)$.
Let us check the stronger statement that the left action defines an
isomorphism $\mathcal{A}_U\cong\Endd_{\,\mhyph\mathcal{A}}(\mathcal{H}_U)$. Recalling how $\mathcal{H}_U$ was
constructed using complementary projections in $\mathcal{A}_{\mathds{1}\oplus U}$, it
becomes clear that the map $\mathcal{A}_U\to\Endd_{\,\mhyph\mathcal{A}}(\mathcal{H}_U)$ is always
surjective, and it is injective if and only if the projection
$p_\mathds{1}\in\mathcal{A}_{\mathds{1}\oplus U}$ has central support $1$. Using the
Frobenius reciprocity isomorphism
$$
\mathcal{C}(V^{\otimes n}\otimes U)\cong \mathcal{C}(V^{\otimes n},V^{\otimes
n}\otimes U\otimes\bar U),
$$
it is easy to check that $\mathcal{H}_{U\otimes\bar U}\cong L^2(\mathcal{A}_U,\tau_U)$
as a Hilbert $\mathcal{A}_U$-$\mathcal{A}$-module. Hence the representation of~$\mathcal{A}_U$ on
$\mathcal{H}_{U\otimes\bar U}$ is faithful. Since $\mathcal{H}_{U\otimes\bar U}\cong
\mathcal{H}_U\otimes_\mathcal{A}\mathcal{H}_{\bar U}$, it follows that the representation of
$\mathcal{A}_U$ on~$\mathcal{H}_U$ is faithful as well. \end{proof}
A similar result could also be proved using the algebras $\mathcal{N}_U$ from
Section~\ref{sec:longo-roberts-appr} instead of $\mathcal{A}_U$. The situation
would be marginally more complicated, since in dealing with the Connes
fusion tensor product~$\otimes_\mathcal{N}$ we would have to take into
account the modular group of $\omega_U$. We are not going to pursue
this topic here, although it could provide a somewhat alternative
route to Proposition~\ref{pminindex} below.
\bigskip
\section{A universal property of the Poisson
boundary}\label{suniversal}
Let $\mathcal{C}$ be a weakly amenable strict C$^*$-tensor category. Fix an
ergodic probability measure $\mu$ on~$\Irr(\mathcal{C})$. Recall that such a
measure exists by Proposition~\ref{pweakamen}. Let
$\Pi\colon\mathcal{C}\to\mathcal{P}$ be the Poisson boundary of~$(\mathcal{C},\mu)$.
For an object $U$ in $\mathcal{C}$ define
$$
d_{\mathrm{min}}^\mathcal{C}(U)=\inf d^\mathcal{A}(F(U)),
$$
where the infimum is taken over all unitary tensor functors
$F\colon\mathcal{C}\to\mathcal{A}$. We will show in the next section that $d_{\mathrm{min}}^\mathcal{C}$
is the amenable dimension function on $\mathcal{C}$. The goal of the present
section is to prove the following.
\begin{theorem} \label{tuniversal1} The Poisson boundary
$\Pi\colon\mathcal{C}\to\mathcal{P}$ is a universal unitary tensor functor such that
$d_{\mathrm{min}}^\mathcal{C}=d^\mathcal{P}\Pi$.
\end{theorem}
In other words, $d_{\mathrm{min}}^\mathcal{C}=d^\mathcal{P}\Pi$ and for any unitary tensor
functor $F\colon\mathcal{C}\to\mathcal{A}$ such that $d_{\mathrm{min}}^\mathcal{C}=d^\mathcal{A} F$ there exists a
unique, up to a natural unitary monoidal isomorphism, unitary tensor
functor $\Lambda\colon\mathcal{P}\to\mathcal{A}$ such that $\Lambda\Pi\cong F$.
\smallskip
Consider a unitary tensor functor $F\colon\mathcal{C}\to\mathcal{A}$, with no
restriction on the dimension function. As we discussed in
Section~\ref{sstensorfunctors}, we may assume that $\mathcal{A}$ is strict,
$\mathcal{C}$ is a C$^*$-tensor subcategory of $\mathcal{A}$ and $F$ is the embedding
functor. Motivated by Izumi's Poisson integral~\cite{MR1916370} we
will define linear maps
$$
\Theta_{U,V}\colon\mathcal{A}(U,V)\to\mathcal{P}(U,V).
$$
We will write $\Theta_U$ for $\Theta_{U,U}$ and often omit the
subscripts altogether, if there is no danger of confusion. The proof
of the theorem will be based on analysis of the multiplicative domain
of $\Theta$.
\smallskip
For every object $U$ in $\mathcal{C}$ fix a standard solution $(R_U,\bar R_U)$
of the conjugate equations in $\mathcal{C}$. Define a faithful state $\psi_U$
on $\End{\mathcal{A}}{U}$ by
$$
\psi_U(T)=||\bar R_U||^{-2}\bar R_U^*(T\otimes\iota)\bar R_U.
$$
Since any other standard solution has the form
$((u\otimes\iota)R_U,(\iota\otimes u)\bar R_U)$ for a unitary $u$,
this definition is independent of any choices. More generally, we can
define in a similar way ``slice maps"
$$
\iota\otimes\psi_V\colon \End{\mathcal{A}}{U\otimes V}\to \End{\mathcal{A}}{U}.
$$
Then, since $((\iota\otimes R_U\otimes \iota)R_V,(\iota\otimes \bar
R_V\otimes \iota)\bar R_U)$ is a standard solution for $U\otimes V$,
we get
\begin{equation} \label{eslice1} \psi_{U\otimes V}=\psi_U(\iota\otimes
\psi_V).
\end{equation}
By definition the state $\psi_U$ extends the trace $\tr_U$ on
$\End{\mathcal{C}}{U}$.
\begin{lemma} The subalgebra $\End{\mathcal{C}}{U}\subset \End{\mathcal{A}}{U}$ is
contained in the centralizer of the state $\psi_U$.
\end{lemma}
\begin{proof} If $u$ is a unitary in $\mathcal{C}(U)$, then the state $\psi_U(u\cdot
u^*)$ is defined similarly to $\psi_U$, but using the solution
$((\iota\otimes u^*)R_U,(u^*\otimes\iota)\bar R_U)$ of the conjugate
equations for $U$. Since $\psi_U$ is independent of the choice of
standard solutions, it follows that $\psi_U(u\cdot u^*)=\psi_U$. But
this exactly means that $\mathcal{C}(U)$ is contained in the centralizer of
$\psi_U$. \end{proof}
It follows that there exists a unique $\psi_U$-preserving conditional
expectation $E_U\colon \End{\mathcal{A}}{U}\to \End{\mathcal{C}}{U}$.
For objects $U$ and $V$ we can consider $\mathcal{A}(U,V)$ as a subspace of
$\End{\mathcal{A}}{U\oplus V}$. Then $E_{U\oplus V}$ defines a linear map
$$
E_{U,V}\colon\mathcal{A}(U,V)\to\mathcal{C}(U,V).
$$
Again, we omit the subscripts when convenient.
\begin{lemma} \label{lcondexp1} The maps $E_{U,V}$ satisfy the
following properties:
\begin{enumerate}
\item $E_{U,V}(T)^*=E_{V,U}(T^*)$;
\item if $T\in\mathcal{A}(U,V)$ and $S\in\mathcal{C}(V,W)$, then
$E_{U,W}(ST)=SE_{U,V}(T)$;
\item for any object $X$ in $\mathcal{C}$ we have $E_{U\otimes
X,V\otimes X}(T\otimes\iota_X)=E_{U,V}(T)\otimes\iota_X$.
\end{enumerate}
\end{lemma}
\begin{proof} Properties (i) and (ii) follows immediately from the corresponding
properties of conditional expectations. To prove (iii), it suffices to
consider the case $U=V$. Take $S\in\End{\mathcal{C}}{U\otimes X}$. Then we
have to check that
$$
\psi_{U\otimes X}(S(T\otimes\iota))=\psi_{U\otimes X}(S(E(T)\otimes
\iota)).
$$
This follows from \eqref{eslice1} and the fact that by definition we
have $(\iota\otimes\psi_X)(S)\in\End{\mathcal{C}}{U}$. \end{proof}
Now, given a morphism $T\in \mathcal{A}(U,V)$, define a bounded natural
transformation $\Theta_{U,V}(T)\colon\iota\otimes U\to\iota\otimes V$
of functors on $\mathcal{C}$ by
$$
\Theta_{U,V}(T)_X=E_{X\otimes U,X\otimes V}(\iota_X\otimes T).
$$
\begin{lemma}\label{luniharmonic} The natural transformation
$\Theta_{U,V}(T)$ is $P_X$-harmonic for any object $X$ in $\mathcal{C}$.
\end{lemma}
\begin{proof} It suffices to consider the case $U=V$. We claim that
$$
(\tr_X\otimes \iota)E(\iota_X\otimes T)=E(T).
$$
Indeed, for any $S\in \End{\mathcal{C}}{U}$ we have
$$
\tr_U\big(S(\tr_X\otimes \iota)E(\iota_X\otimes T)\big) =\tr_{X\otimes
U}(E(\iota_X\otimes ST))\\ =\psi_{X\otimes U}(\iota\otimes ST)
=\psi_U(ST)=\tr_U(SE(T)),
$$
where in the third equality we used \eqref{eslice1}. This proves the
claim.
We now compute:
$$
P_X(\Theta(T))_Y=(\tr_X\otimes\iota)(\Theta(T)_{X\otimes
Y})=(\tr_X\otimes\iota_{Y}\otimes\iota_U)(E(\iota_{X}\otimes\iota_Y\otimes
T))\\ =E(\iota_Y\otimes T)=\Theta(T)_Y,
$$
so $\Theta(T)$ is $P_X$-harmonic. \end{proof}
It follows that $\Theta_{U,V}$ is a well-defined linear map
$\mathcal{A}(U,V)\to \mathcal{P}(U,V)$.
\begin{lemma} \label{ltheta} The maps $\Theta_{U,V}$ satisfy the
following properties:
\begin{itemize}
\item[{\rm (i)}] $\Theta_{U,V}(T)^*=\Theta_{V,U}(T^*)$;
\item[{\rm (ii)}] if $T\in\mathcal{A}(U,V)$ and $S\in\mathcal{C}(V,W)$, then
$\Theta_{U,W}(ST)=S\Theta_{U,V}(T)$;
\item[{\rm (iii)}] for any object $X$ in $\mathcal{C}$ we have
$\Theta_{U\otimes X,V\otimes
X}(T\otimes\iota_X)=\Theta_{U,V}(T)\otimes\iota_X$ and
$\Theta_{X\otimes U,X\otimes V}(\iota_X\otimes T)=\iota_X\otimes
\Theta_{U,V}(T)$;
\item[{\rm (iv)}] the maps $\Theta_U\colon\End{\mathcal{A}}{U}\to\End{\mathcal{P}}{U}$
are unital, completely positive and faithful.
\end{itemize}
\end{lemma}
\begin{proof} All these properties are immediate consequences of the definitions
and the properties of the maps $E_{U,V}$ given in
Lemma~\ref{lcondexp1}. We would like only to point out that the
property $\Theta(\iota\otimes T)=\iota\otimes\Theta(T)$ follows from
the definition of the tensor product in $\mathcal{P}$, the corresponding
property for the maps $E$ is neither satisfied nor needed. \end{proof}
Our goal now is to understand the multiplicative domains of the maps
$\Theta_U\colon\End{\mathcal{A}}{U}\to\End{\mathcal{P}}{U}$. We will first show that
these domains cannot be very large. More precisely, assume we have an
intermediate C$^*$-tensor category $\mathcal{C}\subset\mathcal{B}\subset\mathcal{A}$ such that
$d^\mathcal{A}=d^\mathcal{B}$ on $\mathcal{C}$. For an object $U$ in $\mathcal{C}$ denote by
$E^\mathcal{B}_U\colon\End{\mathcal{A}}{U}\to\mathcal{B}(U)$ the conditional expectation
preserving the categorical trace on $\mathcal{A}$. Then we have the following
result inspired by \cite{MR2335776}*{Lemma~4.5}.
\begin{lemma} \label{lthetafactor} We have $\Theta_U=\Theta_U
E^\mathcal{B}_U$.
\end{lemma}
\begin{proof} We will first show that a similar property holds for the maps $E$,
so $E_U=E_UE^\mathcal{B}_U$.
Consider the normalized categorical trace $\tr^\mathcal{A}_U$ on
$\End{\mathcal{A}}{U}$. We have $\psi_U=\tr^\mathcal{A}_U(\cdot\, Q)$ for some
$Q\in\End{\mathcal{A}}{U}$. The identity $E_U=E_UE^\mathcal{B}_U$ holds if and only if
the conditional expectation $E^\mathcal{B}_U$ is $\psi_U$-preserving, or
equivalently, $Q\in\mathcal{B}(U)$.
By assumption we have $d^\mathcal{A}(U)=d^\mathcal{B}(U)$ for every object $U$ in
$\mathcal{C}$. It follows that a standard solution $(R^\mathcal{B}_U,\bar R^\mathcal{B}_U)$ of
the conjugate equations for $U$ and $\bar U$ in $\mathcal{B}$ remains standard
in $\mathcal{A}$. We have $\bar R_U=(T\otimes\iota)\bar R^\mathcal{B}_U$ for a uniquely
defined $T\in\mathcal{B}(U)$. Then $Q=\frac{d_\mathcal{B}(U)}{d_\mathcal{C}(U)}TT^*\in\mathcal{B}(U)$.
\smallskip
We also need the simple property $E^\mathcal{B}_{X\otimes U}(\iota_X\otimes
T)=\iota_X\otimes E^\mathcal{B}_U(T)$. This is proved similarly to
Lemma~\ref{lcondexp1}(iii), using that $\tr^\mathcal{A}_{X\otimes
U}=\tr^\mathcal{A}_U(\tr^\mathcal{A}_X\otimes\iota)$ and the fact that $\tr^\mathcal{A}$ is
defined using standard solutions in $\mathcal{B}$, so that
$(\tr^\mathcal{A}_X\otimes\iota)(\mathcal{B}(X\otimes U))\subset\mathcal{B}(U)$.
\smallskip
The equality $\Theta_U E^\mathcal{B}_U=\Theta_U$ is now immediate:
$$
\Theta E^\mathcal{B}(T)_X=E(\iota_X\otimes E^\mathcal{B}(T))=E E^\mathcal{B}(\iota_X\otimes T)
=E(\iota_X\otimes T)=\Theta(T)_X.
$$
This proves the assertion.
\end{proof}
Since the completely positive map $\Theta_U$ is faithful, the
multiplicative domain of $\Theta_U=\Theta_UE^\mathcal{B}_U$ is contained in
that of $E^\mathcal{B}_U$, which is exactly $\mathcal{B}(U)$. Therefore to find this
domain we have to consider the smallest possible subcategory that
contains $\mathcal{C}$ and still defines the same dimension function as~$\mathcal{A}$.
\begin{lemma} \label{luniversalau} For every object $U$ in $\mathcal{C}$ there
exists a unique positive invertible element $a_U\in{\mathcal{A}}(U)$ such that
$$
(\iota\otimes a_U^{1/2})R_U\ \ \text{and}\ \
(a^{-1/2}_U\otimes\iota)\bar R_U
$$
form a standard solution of the conjugate equations for $U$ in $\mathcal{A}$.
\end{lemma}
\begin{proof} We can find an invertible element $T\in\End{\mathcal{A}}{U}$ such that
$(\iota\otimes T)R_U$ and $((T^*)^{-1}\otimes\iota)\bar R_U$ form a
standard solution in $\mathcal{A}$. Then we can take $a_U=T^*T$, since
$Ta_U^{-1/2}$ is unitary and hence the morphisms $(\iota\otimes
a_U^{1/2})R_U$ and $(a^{-1/2}_U\otimes\iota)\bar R_U$ still form a
standard solution.
Any other standard solution for $U$ and $\bar U$ has the form
$(\iota\otimes va_U^{1/2})R_U$, $(va^{-1/2}_U\otimes\iota)\bar R_U$
for a unitary $v\in\End{\mathcal{A}}{U}$. By uniqueness of the polar
decomposition the element $va_U^{1/2}$ is positive only if $v=1$. \end{proof}
Note that if we replace $(R_U,\bar R_U)$ by $((\iota\otimes
u)R_U,(u\otimes\iota)\bar R_U)$ for a unitary $u\in\End{\mathcal{C}}{U}$, then
$a_U$ gets replaced by $ua_Uu^*$.
\begin{lemma} \label{luniversaldim} For every object $U$ in $\mathcal{C}$ we
have $d^\mathcal{P}(U)\le d^\mathcal{A}(U)$, and if the equality holds, then we have
$\Theta_U(a_U)^{-1}=\Theta_U(a_U^{-1})$.
\end{lemma}
\begin{proof} As usual, we omit the subscript $U$ in the computations. Consider
the solution
$$
r=(\iota\otimes \Theta(a)^{1/2})R,\ \ \bar
r=(\Theta(a)^{-1/2}\otimes\iota)\bar R
$$
of the conjugate equations for $U$ in $\mathcal{P}$. Then from the equality
$$
r^*r=R^*(\iota\otimes \Theta(a))R=\Theta(R^*(\iota\otimes a)R),
$$
we have $\|r\|=d^{\mathcal{A}}(U)^{1/2}$. On the other hand, we also have
$$
\bar r^*\bar r=\bar R^*(\Theta(a)^{-1}\otimes \iota)\bar R.
$$
By Jensen's inequality for positive maps and the fact that the
function $t\mapsto t^{-1}$ on $(0,+\infty)$ is operator convex (see
e.g.~\cite{MR2251116}*{B.2}), we have
$\Theta(a)^{-1}\le\Theta(a^{-1})$. Hence we have the estimate
$$
\bar r^*\bar r\le\bar R^*(\Theta(a^{-1})\otimes \iota)\bar
R=\Theta(\bar R^*(a^{-1}\otimes \iota)\bar R),
$$
and we conclude that $\|\bar r\|\le d^{\mathcal{A}}(U)^{1/2}$. Hence
$d^\mathcal{P}(U)\le d^\mathcal{A}(U)$, and if the equality holds, then we have $\|\bar
r\|= d^{\mathcal{A}}(U)^{1/2}$ and
$$
\bar R^*(\Theta(a)^{-1}\otimes \iota)\bar R=\bar
R^*(\Theta(a^{-1})\otimes \iota)\bar R.
$$
Since $T\mapsto \bar R^*(T\otimes \iota)\bar R$ is a faithful positive
linear functional on $\End{\mathcal{P}}{U}$, this is equivalent to
$\Theta(a)^{-1}=\Theta(a^{-1})$. \end{proof}
If we have $d^\mathcal{P}(U)=d^\mathcal{A}(U)$, we can then apply the following general result,
which is surely well-known.
\begin{lemma} \label{lmultdomain} Assume $\theta\colon A\to B$ is a
unital completely positive map of C$^*$-algebras and $a\in A$ is a
positive invertible element such that
$\theta(a)^{-1}=\theta(a^{-1})$. Then $a$ lies in the multiplicative
domain of $\theta$.
\end{lemma}
\begin{proof} It suffices to show that $a^{1/2}$ lies in the multiplicative
domain. This, in turn, is equivalent to the equality
$\theta(a)^{1/2}=\theta(a^{1/2})$.
Using Jensen's inequality and operator convexity of the functions
$t\mapsto -t^{1/2}$ and $t\mapsto t^{-1}$, we have
$$
\theta(a)^{1/2}\ge\theta(a^{1/2}),\ \
\theta(a^{-1})^{1/2}\ge\theta(a^{-1/2})\ \ \text{and}\ \
\theta(a^{-1/2})^{-1}\le\theta(a^{1/2}).
$$
The second and the third inequalities imply
$$
\theta(a^{-1})^{-1/2}\le \theta(a^{1/2}).
$$
Since $\theta(a^{-1})=\theta(a)^{-1}$, this gives $\theta(a)^{1/2}\le
\theta(a^{1/2})$. Hence $\theta(a)^{1/2}=\theta(a^{1/2})$. \end{proof}
To finish the preparation for the proof of Theorem~\ref{tuniversal1}
we consider the maps $\Theta$ for $\mathcal{A}=\mathcal{P}$.
\begin{lemma} \label{lthetaid} The maps
$\Theta_{U,V}\colon\mathcal{P}(U,V)\mapsto\mathcal{P}(U,V)$ defined by the functor
$\Pi\colon\mathcal{C}\to\mathcal{P}$ are the identity maps.
\end{lemma}
\begin{proof} It suffices to consider $U=V$. Take $\eta\in\End{\mathcal{P}}{U}$. Let us
show first that $\Theta(\eta)_\mathds{1}=\eta_\mathds{1}$, that is,
$E(\eta)=\eta_\mathds{1}$. In other words, we have to check that for any
$S\in\End{\mathcal{C}}{U}$ we have
$$
\psi_U(S\eta)=\tr_U(S\eta_\mathds{1}).
$$
This follows immediately by definition, since
$$
(\bar R_U^*(S\eta\otimes\iota)\bar R_U)_\mathds{1}=\bar
R_U^*(S\eta_\mathds{1}\otimes\iota)\bar R_U.
$$
Now, for any object $X$ in $\mathcal{C}$, we have
$$
\Theta(\eta)_X=E(\iota_X\otimes\eta)=\Theta(\iota_X\otimes\eta)_\mathds{1}=(\iota_X\otimes\eta)_\mathds{1}=\eta_X.
$$
Therefore we have $\Theta(\eta)=\eta$. \end{proof}
\begin{proof}[Proof of Theorem~\ref{tuniversal1}] The equality
$d_{\mathrm{min}}^\mathcal{C}(U)=d^\mathcal{P}(U)$ for objects $U$ in $\mathcal{C}$ follows from
Lemma~\ref{luniversaldim}.
\smallskip
Let $F\colon\mathcal{C}\to\mathcal{A}$ be a unitary tensor functor such that
$d_{\mathrm{min}}^\mathcal{C}=d^\mathcal{A} F$. As above, we assume that $F$ is simply an
embedding functor. Consider the minimal subcategory
$\tilde\mathcal{B}\subset\mathcal{A}$ containing $\mathcal{C}\subset\mathcal{A}$ and the
morphisms~$\iota_V\otimes a_U\otimes\iota_W$ for all objects $V$, $U$
and $W$ in $\mathcal{C}$, where $a_U\in\End{\mathcal{A}}{U}$ are the morphisms defined
in Lemma~\ref{luniversalau}. This is a C$^*$-tensor subcategory, in
general without subobjects. Complete~$\tilde\mathcal{B}$ with respect to
subobjects to get a C$^*$-tensor category~$\mathcal{B}$. By adding more
objects to $\mathcal{A}$ we may assume without loss of generality that
$\mathcal{B}\subset\mathcal{A}$. Lemmas~\ref{ltheta},~\ref{luniversaldim}
and~\ref{lmultdomain} imply that the maps $\Theta_{U,V}$ define a
strict unitary tensor functor $\tilde\mathcal{B}\to\mathcal{P}$. Thus $\tilde\mathcal{B}$ is
unitary monoidally equivalent to a C$^*$-tensor subcategory
$\tilde\mathcal{P}\subset \mathcal{P}$, possibly without subobjects. Completing
$\tilde\mathcal{P}$ with respect to subobjects we get a C$^*$-tensor
subcategory $\mathcal{P}'\subset\mathcal{P}$, which is unitarily monoidally equivalent
to $\mathcal{B}$.
We claim that the embedding functor $\mathcal{P}'\to\mathcal{P}$ is a unitary monoidal
equivalence. Indeed, by construction we have
$d^{\mathcal{P}'}(U)=d_{\mathrm{min}}^\mathcal{C}(U)$ for every object $U$ in $\mathcal{C}$. By
Lemmas~\ref{lthetafactor} and~\ref{lthetaid} it follows then that the
identity maps $\End{\mathcal{P}}{U}\to\End{\mathcal{P}}{U}$ factor through the
conditional expectations
$E^{\mathcal{P}'}_U\colon\End{\mathcal{P}}{U}\to\End{\mathcal{P}'}{U}$. Hence
$\End{\mathcal{P}}{U}=\End{\mathcal{P}'}{U}$. Since the objects of $\mathcal{C}$ generate
$\mathcal{P}$, this implies that the embedding functor $\mathcal{P}'\to\mathcal{P}$ is a
unitary monoidal equivalence.
We have therefore shown that $\mathcal{P}$ and $\mathcal{B}$ are unitarily monoidally
equivalent, and furthermore, by properties of the maps $\Theta$ such
an equivalence $\Lambda\colon\mathcal{P}\to\mathcal{B}$ can be chosen to be the
identity tensor functor on $\mathcal{C}$. Considered as a functor $\mathcal{P}\to\mathcal{A}$,
the unitary tensor functor $\Lambda$ gives the required factorization
of $F\colon\mathcal{C}\to\mathcal{A}$.
\smallskip
It remains to prove uniqueness. Denote by $\rho_U\in\End{\mathcal{P}}{U}$ the
elements $a_U$ constructed in Lemma~\ref{luniversalau} for the
category $\mathcal{P}$. By the uniqueness part of that lemma, it is clear that
any unitary tensor functor $\Lambda\colon\mathcal{P}\to\mathcal{A}$ extending the
embedding functor $\mathcal{C}\to\mathcal{A}$ must map $\rho_U\in\End{\mathcal{P}}{U}$ into
$a_U\in\End{\mathcal{A}}{U}$. But this completely determines $\Lambda$ up to a
unitary monoidal equivalence, since by the above considerations the
category $\mathcal{P}$ is obtained from $\mathcal{C}$ by adding the morphisms $\rho_U$
and then completing the new category with respect to subobjects. \end{proof}
We finish the section with a couple of corollaries.
\smallskip
The universality of the Poisson boundary implies that up to an
isomorphism the boundary does not depend on the choice of an ergodic
measure. But the proof shows that a stronger result is true.
\begin{corollary} \label{cindepend} Let $\mathcal{C}$ be a weakly amenable
C$^*$-tensor category and $\mu$ be an ergodic probability measure on
$\Irr(\mathcal{C})$. Then any bounded $P_\mu$-harmonic natural transformation
is $P_s$-harmonic for every $s\in\Irr(\mathcal{C})$, so the Poisson boundary
$\Pi\colon\mathcal{C}\to\mathcal{P}$ of $(\mathcal{C},\mu)$ does not depend on the choice of
an ergodic measure.
\end{corollary}
\begin{proof} By Lemma~\ref{lthetaid} the maps
$\Theta_{U,V}\colon\mathcal{P}(U,V)\to\mathcal{P}(U,V)$ are the identity maps, while
by Lem\-ma~\ref{luniharmonic} their images consist of elements that
are $P_s$-harmonic for all $s$. \end{proof}
When $\mathcal{C}$ is amenable, then $d_{\mathrm{min}}^\mathcal{C}=d^\mathcal{C}$ and we get the
following.
\begin{corollary} \label{camenb} If $\mathcal{C}$ is an amenable C$^*$-tensor
category, then its Poisson boundary with respect to any ergodic
probability measure on $\Irr(\mathcal{C})$ is trivial. In other words, any
bounded natural transformation $\iota\otimes U\to\iota\otimes V$ which
is $P_s$-harmonic for all $s\in\Irr(\mathcal{C})$, is defined by a morphism
in~$\mathcal{C}(U,V)$.
\end{corollary}
\begin{proof} The identity functor $\mathcal{C}\to\mathcal{C}$ is already universal, so it is
isomorphic to the Poisson boundary. \end{proof}
We remark that if we were interested only in proving this corollary, a
majority of the above arguments, being applied to the functor
$\mathcal{C}\to\mathcal{P}$, would become either trivial or unnecessary. Namely, in
this case a standard solution of the conjugate equations in~$\mathcal{C}$
remains standard in $\mathcal{P}$, so we have $E_U=E^\mathcal{C}_U$, and the key parts
of the proof are contained in Lemmas~\ref{lthetafactor}
and~\ref{lthetaid}. The first lemma shows that given
$\eta\in\End{\mathcal{P}}{U}$ we have $E(\iota_X\otimes\eta)=E(\iota_X\otimes
E(\eta))$, while the second shows that
$E(\iota_X\otimes\eta)=\eta_X$. Since $E(\iota_X\otimes
E(\eta))=\iota_X\otimes E(\eta)$, we therefore see that $\eta$
coincides with $E(\eta)\in\End{\mathcal{C}}{U}$.
\smallskip
Corollary~\ref{camenb} is more or less known: in view of
Proposition~\ref{prop:PP-equals-HY}, for measures considered
in~\cite{MR1749868} it is equivalent
to~\cite{MR1749868}*{Theorem~7.6}. For an even more restrictive class
of measures the result also follows
from~\cite{MR1444286}*{Theorem~5.16}.
\bigskip
\section{Amenability of the minimal dimension function}
\label{sec:amen-minim-dimens}
As in the previous section, let $\mathcal{C}$ be a weakly amenable strict
C$^*$-tensor category. We defined the dimension function $d_{\mathrm{min}}^\mathcal{C}$
on $\mathcal{C}$ as the infimum of dimension functions under all possible
embeddings of $\mathcal{C}$ into C$^*$-tensor categories, and showed that it
is indeed a dimension function realized by the Poisson boundary of
$\mathcal{C}$ with respect to any ergodic measure. The goal of this section is
to prove the following.
\begin{theorem} \label{tminimal} The dimension function $d_{\mathrm{min}}^\mathcal{C}$ is
amenable, that is, $d_{\mathrm{min}}^\mathcal{C}(U)=\|\Gamma_U\|$ holds for every
object $U$ in $\mathcal{C}$.
\end{theorem}
We remark that already the fact the fusion algebra of a weakly
amenable C$^*$-tensor category admits an amenable dimension function
is nontrivial. We do not know whether this is true for weakly amenable
dimension functions on fusion algebras that are not of categorical
origin. If the fusion algebra is commutative, this is true by a result
of Yamagami~\cite{MR1721584}.
\smallskip
Let $\mu$ be an ergodic probability measure $\mu$ on $\Irr(\mathcal{C})$ and
consider the corresponding Poisson boundary $\Pi\colon\mathcal{C}\to\mathcal{P}$. By
Theorem~\ref{tuniversal1} we already know that
$d_{\mathrm{min}}^\mathcal{C}=d^\mathcal{P}\Pi$. Therefore Theorem~\ref{tminimal} is equivalent
to saying that $d^\mathcal{P}\Pi$ is the amenable dimension function on
$\mathcal{C}$.
\smallskip
We will use the realization of harmonic transformations as elements of
$\mathcal{N}'\cap\mathcal{N}_U$ given in Section~\ref{sec:longo-roberts-appr}. It
will also be important to work with factors. Therefore we assume that
in addition to being ergodic the measure $\mu$ is generating and
satisfies the assumption of
Proposition~\ref{prop:p-bdry-as-rel-comm-LR} (recall that for the
latter it suffices to require $\mu(e)>0$). Recall once again that by
Proposition~\ref{pweakamen} such a measure exists. We also remind that
by Corollary~\ref{cindepend} the Poisson boundary does not depend on
the ergodic measure, but its realization in terms of relative
commutants does. We then have the following expected (in view of
Proposition~\ref{pHYrealization} and the discussion following it), but
crucial, result.
\begin{proposition} \label{pminindex} For every object $U$ in $\mathcal{C}$,
we have $d^\mathcal{P}(U)=[\mathcal{N}_U\colon\mathcal{N}]_0^{1/2}$, where
$[\mathcal{N}_U\colon\mathcal{N}]_0$ is the minimal index of the subfactor
$\mathcal{N}\subset\mathcal{N}_U$.
\end{proposition}
Before we turn to the proof, recall the construction of
$\mathcal{N}_U$. Consider $V=\oplus_{s\in\supp\mu}U_s$. We will work with $V$
as with a well-defined object. If $\supp\mu$ is infinite, to be
rigorous, in what follows we have to replace $V$ by finite sums of
objects $U_s$, $s\in\supp\mu$, and then pass to the limit, but we will
omit this repetitive simple argument. With this understanding,
$\mathcal{N}_U$ is the inductive limit of the algebras
$\mathcal{N}^{(n)}_U=\mathcal{C}(V^{\otimes n}\otimes U)$ equipped with the faithful
states $\omega^{(n)}_U$.
Given another object $U'$, the partial trace $\iota \otimes \tr_{U}$
defines, for each $n$, a conditional expectation $\mathcal{N}_{U' \otimes
U}^{(n)} \to \mathcal{N}_{U'}^{(n)}$ which preserves the state
$\omega^{(n)}_{U'\otimes U}$. The conditional expectation $\mathcal{N}_{U'
\otimes U} \to \mathcal{N}_{U'}$ which we get in the limit, is denoted by
$E_{U',U}$, or simply by $E_U$ if there is no danger of confusion. Fix
a standard solution $(R_U,\bar R_U)$ of the conjugate equations for
$U$ in $\mathcal{C}$.
\begin{lemma}\label{l:basic-ext} The index of the conditional
expectation $E_U\colon\mathcal{N}_U\to\mathcal{N}$ equals $d^\mathcal{C}(U)^2$, the
corresponding basic extension is $\mathcal{N}_U\subset\mathcal{N}_{U \otimes
\bar{U}}$, with the Jones projection $e_U= d^\mathcal{C}(U)^{-1} \bar{R}_U
\bar{R}_U^* \in\mathcal{N}^{(0)}_{U\otimes\bar U}\subset \mathcal{N}_{U \otimes
\bar{U}}$ and the conditional expectation
$E_{\bar{U}}\colon\mathcal{N}_{U\otimes\bar U}\to\mathcal{N}_U$.
\end{lemma}
\begin{proof} By the abstract characterization of the basic extension
\cite{MR1245827}*{Theorem~8} it suffices to check the following three
properties: $E_{\bar U}(e_U)=d^\mathcal{C}(U)^{-2}1$, $E_{\bar
U}(xe_U)e_U=d^\mathcal{C}(U)^{-2}xe_U$ for all $x\in\mathcal{N}_{U\otimes\bar U}$,
and $e_Uxe_U=E_U(x)e_U$ for all $x\in\mathcal{N}_U$. The first and the third
properties are immediate by definition. To prove the second, it is
enough to show that for all $x\in\mathcal{C}(X\otimes U\otimes\bar U)$ we have
$$
d^\mathcal{C}(U)\big((\iota\otimes\tr_{\bar U})(x(\iota_X\otimes \bar R_U\bar
R^*_U))\otimes\iota_{\bar U}\big) (\iota_X\otimes\bar R_U\bar
R_U^*)=x(\iota_X\otimes\bar R_U\bar R_U^*).
$$
The left hand side equals
\begin{gather*}
(\iota_{X}\otimes\iota_{U}\otimes R^*_U\otimes\iota_{\bar U})
(x\otimes\iota_{U}\otimes\iota_{\bar U}) (\iota_X\otimes \bar
R_U\bar R^*_U\otimes\iota_{U}\otimes\iota_{\bar U})
(\iota_{X}\otimes\iota_{U}\otimes R_U\otimes\iota_{\bar U})
(\iota_X\otimes\bar R_U\bar R_U^*)\\
\begin{split}
&= (\iota_{X}\otimes\iota_{U}\otimes R^*_U\otimes\iota_{\bar U})
(x\otimes\iota_{U}\otimes\iota_{\bar U}) (\iota_X\otimes \bar
R_U\otimes\iota_{U}\otimes\iota_{\bar U}) (\iota_X\otimes\bar R_U\bar R_U^*)\\
&= (\iota_{X}\otimes\iota_{U}\otimes R^*_U\otimes\iota_{\bar U})
(x\otimes\iota_{U}\otimes\iota_{\bar U}) (\iota_X\otimes \bar
R_U\otimes\bar R_U) (\iota_X\otimes\bar R_U^*)\\
&=x(\iota_X\otimes\bar R_U\bar R_U^*),
\end{split}
\end{gather*}
which proves the lemma.
\end{proof}
This lemma implies in particular that there exists a unique
representation $$\pi\colon\mathcal{N}_{U\otimes\bar U}\to
B(L^2(\mathcal{N}_U,\omega_U))$$ that extends the representation of $\mathcal{N}_U$
and is such that $\pi(e_U)$ is the projection onto the closure of
$\Lambda_{\omega_U}(\mathcal{N})\subset L^2(\mathcal{N}_U,\omega_U)$.
\begin{lemma} \label{lpi} The representation
$\pi\colon\mathcal{N}_{U\otimes\bar U}\to B(L^2(\mathcal{N}_U,\omega_U))$ is given
by
$$
\pi(x)\Lambda_{\omega_U}(y) = \Lambda_{\omega_U}((\iota\otimes
R_U^*)(x\otimes\iota_U)(y\otimes\iota_{\bar U}\otimes
\iota_U)(\iota\otimes \bar R_U\otimes\iota_U))
$$
for $x\in\cup_n\mathcal{N}^{(n)}_{U\otimes\bar U}$ and
$y\in\cup_n\mathcal{N}^{(n)}_U$.
\end{lemma}
\begin{proof} Let us write $\tilde\pi(x)$ for the operators in the formulation
of the lemma. The origin of the formula for $\tilde\pi$ is the
Frobenius reciprocity isomorphism
$$
\mathcal{C}(V^{\otimes n}\otimes U) \cong \mathcal{C}(V^{\otimes n}, V^{\otimes
n}\otimes U\otimes \bar U),\ \ T\mapsto (T\otimes\iota_{\bar
U})(\iota_{V^{\otimes n}}\otimes\bar R_U),
$$
with inverse $S\mapsto (\iota\otimes R^*_U)(S\otimes\iota_{U})$. Up to
scalar factors these isomorphisms become unitary once we equip both
spaces with scalar products defined by the states $\omega^{(n)}_U$ and
$\omega^{(n)}_\mathds{1}$, respectively. The algebra $\mathcal{C}(V^{\otimes
n}\otimes U\otimes \bar U)$ is represented on $ \mathcal{C}(V^{\otimes
n},V^{\otimes n}\otimes U\otimes \bar U)$ by the operators of
multiplication on the left. Being written on the space $\mathcal{C}(V^{\otimes
n}\otimes U)$, this representation is exactly $\tilde\pi$. Therefore
$\tilde\pi$ certainly defines a representation of the $*$-algebra
$\cup_n\mathcal{N}^{(n)}_{U\otimes\bar U}$ on the dense subspace $\cup_n
L^2(\mathcal{N}^{(n)}_U,\omega^{(n)}_U)$ of $L^2(\mathcal{N}_U,\omega_U)$. In order
to see that this representation extends to a normal representation of
$\mathcal{N}_{U\otimes\bar U}$, observe that the vector
$\Lambda_{\omega_U}(1)$ is cyclic and
$$
\left(\tilde\pi(x) \Lambda_{\omega_U}(1), \Lambda_{\omega_U}(1)\right)
= d^\mathcal{C}(U)^2 \omega_{U\otimes\bar U}(e_{U} x e_{U}),
$$
since for every $z\in\mathcal{C}(U\otimes\bar U)$ we have
\begin{align*} \tr_U((\iota_U\otimes R_U^*)(z\otimes\iota_U)(\bar
R_U\otimes\iota_U)) &=d^\mathcal{C}(U)^{-1}\bar R_U^*(\iota_U\otimes
R_U^*\otimes\iota_{\bar U})(z\otimes\iota_U\otimes\iota_{\bar U})(\bar
R_U\otimes\iota_U\otimes\iota_{\bar U})\bar R_U\\ &=d^\mathcal{C}(U)^{-1}\bar
R_U^*z\bar R_U=d^\mathcal{C}(U)\tr_{U\otimes\bar U}(z\bar R_U^*\bar R_U)\\
&=d^\mathcal{C}(U)^2\tr_{U\otimes\bar U}(ze_U)=d^\mathcal{C}(U)^2\tr_{U\otimes\bar
U}(e_Uze_U).
\end{align*}
It is clear that
$\tilde\pi(x)\Lambda_{\omega_U}(y)=\Lambda_{\omega_U}(xy)$ for
$x\in\cup_n\mathcal{N}^{(n)}_U$, so $\tilde\pi$ extends the representation of
$\mathcal{N}_U$ on $L^2(\mathcal{N}_U,\omega_U)$. Therefore to prove that
$\pi=\tilde\pi$ it remains to show that $\tilde\pi(e_U)$ is the
projection onto $\overline{\Lambda_{\omega_U}(\mathcal{N})}$, that~is,
$$
\tilde\pi(e_U)\Lambda_{\omega_U}(y)=\Lambda_{\omega_U}(E_U(y))\ \
\text{for}\ \ y\in\cup_n\mathcal{N}^{(n)}_U.
$$
But this is obvious, as $(\iota_U\otimes R^*_U)(\bar R_U\bar
R^*_U\otimes\iota_U)=\bar R^*_U\otimes\iota_U$. \end{proof}
It is easy to describe the modular group $\sigma^{\omega_U}$ of
$\omega_U$. For $s_* = (s_1, \ldots, s_n) \in (\supp \mu)^n$, let us
put
$$
\delta_{s_*} = \frac{\mu(s_1) \cdots \mu(s_n)}{d^\mathcal{C}(U_{s_1}) \cdots
d^\mathcal{C}(U_{s_n})}.
$$
Then
$$
\sigma^{\omega_U}_t(x)=\left(\frac{\delta_{s'_*}}{\delta_{s_*}}\right)^{it}x\
\ \text{for}\ \ x\in \mathcal{C}(U_{s_n} \otimes \cdots \otimes U_{s_n}\otimes
U, U_{s'_n} \otimes \cdots \otimes U_{s'_1}\otimes U).
$$
What matters for us is that since the automorphisms
$\sigma^{\omega_U}_t$ are approximately implemented by unitaries in
$\mathcal{N}$, the relative commutant $\mathcal{N}'\cap\mathcal{N}_U$ is contained in the
centralizer of the state $\omega_U$.
\smallskip
Consider the modular conjugation $J=J_{\omega_U}$ on
$L^2(\mathcal{N}_U,\omega_U)$. By Lemma~\ref{l:basic-ext} and definition of
the basic extension we have
$$
J\mathcal{N}'J=\pi(\mathcal{N}_{U\otimes\bar U}).
$$
Therefore the map $x\mapsto Jx^*J$ defines a $*$-anti-isomorphism
$$
\mathcal{N}'\cap\mathcal{N}_U\cong \mathcal{N}_U'\cap\mathcal{N}_{U\otimes\bar U}.
$$
Identifying these relative commutants with $\mathcal{P}(U)$ and $\mathcal{P}(\bar U)$,
respectively, we get a $*$-anti-isomorphism $\mathcal{P}(U)\cong\mathcal{P}(\bar U)$,
which we denote by $\eta\mapsto\eta^\vee$.
\begin{lemma} \label{lvee} For every $\eta\in\mathcal{P}(U)$ we have
$$
\eta^\vee=(R^*_U\otimes\iota_{\bar U})(\iota_{\bar
U}\otimes\eta\otimes\iota_{\bar U})(\iota_{\bar U}\otimes\bar R_U).
$$
\end{lemma}
\begin{proof} Consider the element $\tilde\eta=(R^*_U\otimes\iota_{\bar
U})(\iota_{\bar U}\otimes\eta\otimes\iota_{\bar U})(\iota_{\bar
U}\otimes\bar R_U)$. In terms of families of morphisms this means that
$$
\tilde\eta_X=(\iota_X\otimes R^*_U\otimes\iota_{\bar
U})(\eta_{X\otimes\bar U}\otimes\iota_{\bar
U})(\iota_X\otimes\iota_{\bar U}\otimes\bar R_U),
$$
or equivalently,
\begin{equation}\label{evee} (\iota_X\otimes
R^*_U)(\tilde\eta_X\otimes\iota_U)=(\iota_X\otimes
R^*_U)\eta_{X\otimes\bar U}.
\end{equation}
For every $n$ consider the projection $p_n\colon
L^2(\mathcal{N}_U,\omega_U)\to L^2(\mathcal{N}^{(n)}_U,\omega^{(n)}_U)$. Let
$x\in\mathcal{N}'\cap\mathcal{N}_U$ be the element corresponding to $\eta$, and
$\tilde x\in\mathcal{N}_U'\cap\mathcal{N}_{U\otimes\bar U}$ be the element
corresponding to $\tilde\eta$. By Lemma~\ref{lpi} and the way we
represent $\tilde\eta$ by $\tilde x$, for every $y\in\mathcal{N}^{(n)}_U$ we
have
\begin{equation} \label{evee2} p_n\pi(\tilde
x)\Lambda_{\omega_U}(y)=\Lambda_{\omega_U}((\iota\otimes
R_U^*)(\tilde\eta_{V^{\otimes n}\otimes
U}\otimes\iota_U)(y\otimes\iota_{\bar U}\otimes \iota_U)(\iota\otimes
\bar R_U\otimes\iota_U)).
\end{equation} On the other hand, since $x$ is contained in the
centralizer of $\omega_U$, we have
\begin{align*}
p_nJx^*J\Lambda_{\omega_U}(y)&=p_n\Lambda_{\omega_U}(yx)=\Lambda_{\omega_U}(y
\eta_{V^{\otimes n}})\\ &=\Lambda_{\omega_U}((\iota\otimes
R_U^*)(y\otimes\iota_{\bar U}\otimes \iota_U)(\iota\otimes \bar
R_U\otimes\iota_U)\eta_{V^{\otimes n}})\\
&=\Lambda_{\omega_U}((\iota\otimes R_U^*)\eta_{V^{\otimes n}\otimes
U\otimes\bar U}(y\otimes\iota_{\bar U}\otimes \iota_U)(\iota\otimes
\bar R_U\otimes\iota_U)).
\end{align*} By \eqref{evee} the last expression equals \eqref{evee2},
so
$$
p_n\pi(\tilde x)\Lambda_{\omega_U}(y)=p_nJx^*J\Lambda_{\omega_U}(y).
$$
Since this is true for all $n$ and $y\in\mathcal{N}^{(n)}_U$, we conclude
that $\pi(\tilde x)=Jx^*J$. \end{proof}
\begin{proof}[Proof of Proposition~\ref{pminindex}] The operator valued weights
from $\mathcal{N}_U$ to $\mathcal{N}$ are parametrized by the positive elements $a
\in \mathcal{N}'\cap\mathcal{N}_U$ by $a \mapsto E^a$, where $E^a$ is defined by
$E^a(x) = E_U(a^{1/2} x a^{1/2})$. The map $E^a$ is a conditional
expectation if and only if the normalization condition $E_U(a)=1$
holds. Moreover, by the proof of~\cite{MR976765}*{Theorem~1},
$(E^a)^{-1}$ is given by $x \mapsto E_U^{-1}(a^{-1/2} x
a^{-1/2})$. Therefore we have
$$
[\mathcal{N}_U\colon\mathcal{N}]_0=\min_{\substack{a\in \mathcal{N}'\cap \mathcal{N}_U,\\
a>0}}E_U(a)E^{-1}_U(a^{-1})=\min_{\substack{a\in \mathcal{N}'\cap \mathcal{N}_U,\\
a>0}}d^\mathcal{C}(U)^2E_U(a)\tilde E_U(Ja^{-1}J),
$$
where $\tilde E_U=d^\mathcal{C}(U)^{-2}JE_U^{-1}(J\cdot
J)J\colon\pi(\mathcal{N}_{U\otimes\bar U})\to\mathcal{N}_U$.
If $a\in \mathcal{N}'\cap \mathcal{N}_U$ corresponds to $\eta\in\mathcal{P}(U)$, we have
$$
E_U(a)=d^{\mathcal{C}}(U)^{-1}\bar R_U(\eta\otimes\iota)\bar R_U^*
$$
By Lemma~\ref{l:basic-ext} we have $\tilde E_U(\pi(x))=\pi(E_{\bar
U}(x))$ for $x\in\mathcal{N}_{U\otimes\bar U}$. Hence by Lemma~\ref{lvee} we
get
\begin{align*} \tilde
E_U(Ja^{-1}J)&=d^{\mathcal{C}}(U)^{-1}R_U^*((\eta^{-1})^\vee\otimes\iota)R_U\\
&=d^{\mathcal{C}}(U)^{-1}R_U^*(R^*_U\otimes\iota_{\bar
U}\otimes\iota_U)(\iota_{\bar U}\otimes\eta^{-1}\otimes\iota_{\bar
U}\otimes\iota_U)(\iota_{\bar U}\otimes\bar R_U\otimes\iota_U)R_U\\
&=d^{\mathcal{C}}(U)^{-1}R_U^*(\iota_{\bar U}\otimes\eta^{-1})R_U.
\end{align*} We thus conclude that $[\mathcal{N}_U\colon\mathcal{N}]_0$ is the
minimum of the products of the scalars
$$
\bar R_U(\eta\otimes\iota)\bar R_U^*\ \ \text{and}\ \
R_U^*(\iota\otimes\eta^{-1})R_U
$$
over all positive invertible $\eta\in\mathcal{P}(U)$. This is exactly
$d^\mathcal{P}(U)^2$. \end{proof}
\begin{proof}[Proof of Theorem~\ref{tminimal}] The estimate $\| \Gamma_U \| \le
d^\mathcal{P}(U)$ comes for free. We thus need to prove the opposite
inequality.
Let $E^\mathcal{P}_U\colon\mathcal{N}_U\to\mathcal{N}$ be the minimal conditional
expectation. Let us first assume that $\mathcal{N}_U$ (and hence $\mathcal{N}$) is
infinite. Then by Proposition~\ref{pminindex} and
\cite{MR1096438}*{Corollary~7.2} we have the equalities
$$
2\log d^\mathcal{P}(U)=\log \Ind E^\mathcal{P}_U =H_{E^\mathcal{P}_U}(\mathcal{N}_U | \mathcal{N}).
$$
Let $\epsilon > 0$ and $\psi$ be a normal state on $\mathcal{N}_U$ such that
$$H_{\psi}(\mathcal{N}_U | \mathcal{N}) \ge 2 \log d^\mathcal{P}(U)
- \epsilon.$$
When $A$ is a finite subset of $\supp \mu$, consider the projection
$p_A = \oplus_{s \in A} \iota_s$ in $\mathcal{N}^{(1)}$. If $A_1, \ldots,
A_n$ are finite subsets of $\supp\mu$, then $p_{A_*} = p_{A_n} \otimes
\cdots \otimes p_{A_1}$ is a projection in $\mathcal{N}^{(n)}$, and we
consider the corresponding corner
$$
\mathcal{N}_U^{A_*} = p_{A_*}\mathcal{N}_U^{(n)} p_{A_*} = \bigoplus_{\substack{s_i,
s'_i \in A_i\\i=1,\ldots,n}} \mathcal{C}(U_{s_n} \otimes \cdots \otimes
U_{s_1} \otimes U, U_{s'_n} \otimes \cdots \otimes U_{s'_1} \otimes U)
$$
in $\mathcal{N}^{(n)}_U$ and the similarly defined corner $\mathcal{N}^{A_*}$ in
$\mathcal{N}^{(n)}$. When $\psi(p_{A_*})\ne0$, define also a state
$\psi_{A_*}$ on~$\mathcal{N}_U^{A_*} $ by $
\psi_{A_*}=\psi(p_{A_*})^{-1}\psi(p_{A_*}\cdot p_{A_*}). $ By the
lower semicontinuity of relative entropy, we can find~$n$ and finite
sets $A_1, \ldots, A_n$ such that
$$
H_{\psi_{A_*}}(\mathcal{N}^{A_*}_U | \mathcal{N}^{A_*}) \ge H_\psi(\mathcal{N}_U | \mathcal{N}) -
\epsilon.
$$
By Proposition~\ref{prop:stt-rel-entropy-graph-norm}, the inclusion
matrix $\Gamma_{A_*, U}$ of $\mathcal{N}^{A_*} \subset \mathcal{N}_U^{A_*}$
satisfies $$2 \log \| \Gamma_{A_*,U} \|\ge H_{\psi_{A_*}}(
\mathcal{N}_U^{A_*} | \mathcal{N}^{A_*}).$$ Therefore we have the estimate
$$
\log \| \Gamma_{A_*, U} \|\ge\log d^\mathcal{P}(U)- \epsilon.
$$
But the transpose of the matrix $\Gamma_{A_*,U}$ is obtained from
$\Gamma_U$ by considering only columns that correspond to the simple
objects appearing in the decomposition of $U_{s_n} \otimes \cdots
\otimes U_{s_1} $ for $s_i\in A_i$, and then removing the zero
rows. Hence
$$
\|\Gamma_U \|\ge \| \Gamma_{A_*,U} \|.
$$
Since $\epsilon$ was arbitrary, we thus get $\| \Gamma_U \| \ge
d^\mathcal{P}(U)$.
\smallskip
If $\mathcal{N}_U$ is finite, we consider the inclusion $\mathcal{N} \mathbin{\bar{\otimes}} M
\subset \mathcal{N}_U \mathbin{\bar{\otimes}} M$ for some infinite hyperfinite von Neumann
algebra $M$ with a prescribed strongly operator dense increasing
sequence $M_{n_k}(\mathbb{C})\subset M$; for example, we could take a Powers
factor $R_\lambda$ with the usual copies of $M_2(\mathbb{C})^{\otimes k}$ in
it. Then the minimal conditional expectation $\mathcal{N}_U \mathbin{\bar{\otimes}} M \to
\mathcal{N} \mathbin{\bar{\otimes}} M$ is given by $E^\mathcal{P}_U \otimes \iota$, and its index
equals that of $E^\mathcal{P}_U$. Since the inclusion matrix of $\mathcal{N}^{A_*}
\otimes M_{n_k}(\mathbb{C}) \subset \mathcal{N}_U^{A_*} \otimes M_{n_k}(\mathbb{C})$ is the
same as that of $\mathcal{N}^{A_*} \subset \mathcal{N}_U^{A_*}$, we can then argue
in the same way as above. \end{proof}
Since amenability of dimension functions is preserved under
homomorphisms of fusion algebras by
\cite{MR1644299}*{Proposition~7.4}, we get the following corollary.
\begin{corollary}\label{cor:from-P-eq-C-to-amen} Let
$\Pi\colon\mathcal{C}\to\mathcal{P}$ be the Poisson boundary of a rigid C$^*$-tensor
category with respect to an ergodic probability measure on
$\Irr(\mathcal{C})$. Then $\mathcal{P}$ is an amenable C$^*$-tensor category.
\end{corollary}
Combining this with Corollary~\ref{camenb} we get the following
categorical version of the
Furstenberg--Kaimanovich--Vershik--Rosenblatt characterization of
amenability.
\begin{theorem} \label{tFKVR} A rigid C$^*$-tensor category $\mathcal{C}$ is
amenable if and only if there is a probability measure~$\mu$ on
$\Irr(\mathcal{C})$ such that the Poisson boundary of $(\mathcal{C},\mu)$ is
trivial. Furthermore, the Poisson boundary of an amenable C$^*$-tensor
category is trivial for any ergodic probability measure.
\end{theorem}
Therefore we can say that while weak amenability can be detected by
studying classical Poisson boundaries of random walks on the fusion
algebra, for amenability we have to consider noncommutative, or
categorical, random walks. We can also say that nontriviality of the
Poisson boundary $\Pi\colon\mathcal{C}\to\mathcal{P}$ with respect to an ergodic
measure shows how far a weakly amenable category $\mathcal{C}$ is from being
amenable.
\bigskip
\section{Amenable functors}
In this section we will give another characterization of amenability
in terms of invariant means. We know that on the level of fusion
algebras existence of invariant means is not enough for
amenability. Therefore we need a more refined categorical notion.
\begin{definition} Let $\mathcal{C}$ be a C$^*$-tensor category and
$F\colon\mathcal{C}\to\mathcal{A}$ be a unitary tensor functor into a C$^*$-tensor
category $\mathcal{A}$ with possibly nonsimple unit. A \emph{right invariant
mean} for $F$ is a collection $m=(m_{U,V})_{U,V}$ of linear
maps $$m_{U,V}\colon \hat{\mathcal{C}}(U, V) \to \mathcal{A}(F(U),F(V))$$ that are
natural in $U$ and $V$ and satisfy the following properties:
\begin{enumerate}
\item the maps $m_U=m_{U,U}\colon \hat{\mathcal{C}}(U)\to
\End{\mathcal{A}}{F(U)}$ are unital and positive;
\medskip
\item for any $\eta\in \hat{\mathcal{C}}(U, V)$ and any object $Y$ in
$\mathcal{C}$ we have
$$
m_{U\otimes Y,V\otimes
Y}(\eta\otimes\iota_Y)=F_2(m_{U,V}(\eta)\otimes\iota_{F(Y)});
$$
\item for any $\eta\in \hat{\mathcal{C}}(U, V)$ and any object $Y$ in
$\mathcal{C}$ we have
$$
m_{Y\otimes U,Y\otimes V}(\iota_Y\otimes\eta)=F_2(\iota_{F(Y)}\otimes
m_{U,V}(\eta)).
$$
\end{enumerate}
If a right invariant mean for $F$ exists, we say that $F$ is
\emph{amenable}.
\end{definition}
Note that naturality of $m_{U,V}$ and property (i) in the above
definition easily imply that the maps~$m_U$ are completely positive,
and $m_{U,V}(\eta)^*=m_{V,U}(\eta^*)$. As usual, we omit subscripts
and simply write $m$ instead of $m_{U, V}$ when there is no confusion.
\smallskip
The relevance of this notion for categorical random walks is explained
by the following simple observation, similar to the easy part of
Proposition~\ref{pweakamen}.
\begin{proposition} \label{ppoissonamenability} Let $\mathcal{C}$ be a rigid
C$^*$-tensor category, $\mu$ be a probability measure on $\Irr(\mathcal{C})$,
and $\Pi\colon\mathcal{C}\to\mathcal{P}$ be the Poisson boundary of~$(\mathcal{C},\mu)$. Then
the functor $\Pi\colon\mathcal{C}\to\mathcal{P}$ is amenable.
\end{proposition}
\begin{proof} Fix a free ultrafilter $\omega$ on $\mathbb{N}$, and then define
$$
m(\eta)_X=\lim_{n\to\omega}\frac{1}{n}\sum^{n-1}_{k=0}P^k_\mu(\eta)_X.
$$
All the required properties of a right invariant mean follow
immediately by definition. For example, property (iii) in the
definition follows from the identity $P_X(\iota_Y\otimes
\eta)=\iota_Y\otimes P_X(\eta)$. \end{proof}
For functors into categories with nonsimple units we do not have much
insight into the meaning of amenability. But if we fall back to our
standard assumption of simplicity of tensor units, we have the
following result.
\begin{theorem}\label{tamnefunctor} Let $\mathcal{C}$ be a rigid C$^*$-tensor
category and $F\colon\mathcal{C}\to\mathcal{A}$ be a unitary tensor functor. Then~$F$
is amenable if and only if $\mathcal{C}$ is weakly amenable and $d^\mathcal{A} F$ is
the amenable dimension function on~$\mathcal{C}$.
\end{theorem}
Let $F\colon\mathcal{C}\to\mathcal{A}$ be an amenable unitary tensor functor with a
right invariant mean $m$. For simplicity we assume as usual that $\mathcal{C}$
and $\mathcal{A}$ are strict and $F$ is an embedding functor. Let us start by
showing that existence of $F$ implies weak amenability.
\begin{lemma} \label{lweakamen} The linear functional
$m_\mathds{1}\colon\Endb(\iota_\mathcal{C})\cong\ell^\infty(\Irr(\mathcal{C}))\to\End{\mathcal{A}}{\mathds{1}}\cong\mathbb{C}$
is a right invariant mean on the fusion algebra of $\mathcal{C}$ equipped with
the dimension function $d^\mathcal{C}$.
\end{lemma}
\begin{proof} In addition to the operators $P_X$ on $\hat\mathcal{C}(\mathds{1})$ we
normally use, we also have the operators $Q_X$ given by
$$
Q_X(\eta)_Y=d^\mathcal{C}(X)^{-1}(\iota_Y\otimes\tr_X)(Q_{Y\otimes
X})=d^\mathcal{C}(X)^{-1} (\iota_Y\otimes\bar R_X^*)(\eta_{Y\otimes
X}\otimes\iota_{\bar X})(\iota_Y\otimes \bar R_X),
$$
where $(R_X,\bar R_X)$ is a standard solution of the conjugate
equations for $X$ in $\mathcal{C}$. Since
$$
\eta_{Y\otimes X}\otimes\iota_{\bar
X}=(\iota_X\otimes\eta\otimes\iota_{\bar X})_Y,
$$
we can write this as
$$
Q_X(\eta)=d^\mathcal{C}(X)^{-1}\bar R^*_X(\iota_X\otimes\eta\otimes\iota_{\bar
X})\bar R_X.
$$
Applying the invariant mean we get
$$
m(Q_X(\eta))=d^\mathcal{C}(X)^{-1}\bar R^*_X(\iota_X\otimes
m(\eta)\otimes\iota_{\bar X})\bar R_X=m(\eta).
$$
Thus $m_\mathds{1}$ is a right invariant mean on
$\ell^\infty(\Irr(\mathcal{C}))$. \end{proof}
Since $\mathcal{C}$ is weakly amenable, we can choose an ergodic probability
measure and consider the corresponding Poisson boundary
$\Pi\colon\mathcal{C}\to\mathcal{P}$. We then have the following result, which has its
origin in Tomatsu's considerations in~\cite{MR2335776}*{Section~4}.
\begin{lemma} \label{lmult2} For every object $U$ in $\mathcal{C}$ the map
$\Lambda_U\colon\End{\mathcal{P}}{U}\to\End{\mathcal{A}}{U}$ obtained by
restricting~$m_U$ to~$\End{\mathcal{P}}{U}$ is multiplicative.
\end{lemma}
\begin{proof} Recall that in Section~\ref{suniversal} we constructed faithful
unital completely positive maps
$\Theta_U\colon\End{\mathcal{A}}{U}\to\End{\mathcal{P}}{U}$, $\Theta_U(T)_X=E_{X\otimes
U}(\iota\otimes T)$. By faithfulness of $\Theta_U$, the multiplicative
domain of $\Theta_U\Lambda_U$ is contained in that of
$\Lambda_U$. Therefore in order to prove the lemma it suffices to show
that $\Theta_U\Lambda_U$ is the identity map.
Let us show first that for any $\eta\in\End{\mathcal{P}}{U}$ we have $\Theta
\Lambda(\eta)_\mathds{1}=\eta_\mathds{1}$, that is,
$$
E(m(\eta))=\eta_\mathds{1}.
$$
Take $S\in\End{\mathcal{C}}{U}$. Then we have
$$
\tr_U(SE(m(\eta)))=\psi_U(m(S\eta))=d(U)^{-1}\bar
R_U^*(m(S\eta)\otimes\iota)\bar R_U =d(U)^{-1}m(\bar
R_U^*(S\eta\otimes\iota)\bar R_U).
$$
Since the element $\bar R_U^*(S\eta\otimes\iota)\bar R_U$ lies in
$\End{\mathcal{P}}{\mathds{1}}$, it is scalar. This scalar must be equal to
$$\bar R_U^*(S\eta_\mathds{1}\otimes\iota)\bar R_U=\tr_U(S\eta_\mathds{1}).$$
Hence we obtain
$$
\tr_U(SE(m(\eta)))=\tr_U(S\eta_\mathds{1}),
$$
and since this is true for all $S$, we get $\Theta
\Lambda(\eta)_\mathds{1}=\eta_\mathds{1}$.
Now, for any object $X$ in $\mathcal{C}$, we use the above equality for
$\iota_X\otimes\eta$ instead of $\eta$ and get
$$
\Theta \Lambda(\eta)_X=E(\iota_X\otimes
m(\eta))=E(m(\iota_X\otimes\eta))=\Theta
\Lambda(\iota_X\otimes\eta)_\mathds{1} =(\iota_X\otimes\eta)_\mathds{1}=\eta_X,
$$
which implies the desired equality $\Theta \Lambda(\eta)=\eta$. \end{proof}
\begin{proof}[Proof of Theorem~\ref{tamnefunctor}] Consider an amenable unitary
tensor functor $F\colon\mathcal{C}\to\mathcal{A}$.
By Lemma~\ref{lweakamen} we know that
$\mathcal{C}$ is weakly amenable. By Lemma~\ref{lmult2} and the definition of
invariant means, any right invariant mean for $F$ defines a strict
unitary tensor functor $\Lambda\colon\tilde\mathcal{P}\to\mathcal{A}$, where
$\tilde\mathcal{P}\subset\mathcal{P}$ is the full subcategory consisting of objects in
$\mathcal{C}$. Extend this functor to a unitary tensor functor
$\Lambda\colon\mathcal{P}\to\mathcal{A}$. Then $d^\mathcal{A}(U)\le d^\mathcal{P}(U)$ for any object $U$
in $\mathcal{C}$, but since by Theorem~\ref{tminimal} the dimension function
$d^\mathcal{P}\Pi$ on $\mathcal{C}$ is amenable, we conclude that
$d^\mathcal{A}(U)=d^\mathcal{P}(U)=\|\Gamma_U\|$.
\smallskip
Conversely, assume $\mathcal{C}$ is weakly amenable and $F\colon\mathcal{C}\to\mathcal{A}$ is a
unitary tensor functor such that $d^\mathcal{A} F$ is the amenable dimension
function. Then by Theorem~\ref{tuniversal1} there exists a unitary
tensor functor $\Lambda\colon\mathcal{P}\to\mathcal{A}$ such that $\Lambda\Pi\cong
F$. By Proposition~\ref{ppoissonamenability} there exists a right
invariant mean for the functor $\Pi\colon\mathcal{C}\to\mathcal{P}$. Composing it with
the functor $\Lambda$ we get a right invariant mean for $\Lambda\Pi$,
from which we get a right invariant mean for $F$. \end{proof}
Applying Theorem~\ref{tamnefunctor} to the identity functor we get a
characterization of amenability of tensor categories in terms of
invariant means.
\begin{theorem} A rigid C$^*$-tensor category $\mathcal{C}$ is amenable if and
only if the identity functor $\mathcal{C}\to\mathcal{C}$ is amenable.
\end{theorem}
Note that by the proof of Theorem~\ref{tamnefunctor}, given an
amenable C$^*$-tensor category $\mathcal{C}$, we can construct a right
invariant mean for the identity functor as follows. Choose an ergodic
probability measure $\mu$ on $\Irr(\mathcal{C})$ and a free ultrafilter
$\omega$ on $\mathbb{N}$. Then we can define
$$
m(\eta)=\lim_{n\to\omega}\frac{1}{n}\sum^{n-1}_{k=0}
P^k_\mu(\eta)_\mathds{1}.
$$
On the other hand, the construction of a right invariant mean for a
functor $F\colon\mathcal{C}\to\mathcal{A}$ such that $\mathcal{C}$ is weakly amenable, but not
amenable, and $d^\mathcal{A} F$ is the amenable dimension function, is more
elusive, as it relies on the existence of a factorization of $F$
through the Poisson boundary $\mathcal{C}\to\mathcal{P}$.
\bigskip
\section{Amenability of quantum groups and subfactors}
In this section we apply some of our results to categories considered
in the theory of compact quantum groups and subfactor theory.
\subsection{Quantum groups}
Let $G$ be a compact quantum group. We follow the conventions
of~\cite{arXiv:1310.4407}. In particular, the algebra $\mathbb{C}[G]$ of
regular functions on $G$ is a Hopf $*$-algebra, and by a finite
dimensional unitary representation of $G$ we mean a unitary element
$U\in B(H_U)\otimes\mathbb{C}[G]$, where $H_U$ is a finite dimensional Hilbert
space, such that $(\iota\otimes\Delta)(U)=U_{12}U_{13}$. Finite
dimensional unitary representations form a rigid C$^*$-tensor category
$\Rep G$, with the tensor product of $U$ and $V$ defined by
$U_{13}V_{23}\in B(H_U)\otimes B(H_V)\otimes\mathbb{C}[G]$. The categorical
dimension of $U$ is equal to the quantum dimension, given by the trace $\Tr(\rho_U)$ of the Woronowicz character.
\smallskip
From the universal property of the Poisson boundary it is easy to
deduce that the Poisson boundary of $\Rep G$ with respect to any
ergodic measure is the forgetful functor $\Rep G\to \Rep K$, where
$K\subset G$ is the maximal quantum subgroup of $G$ of Kac type. This
will be discussed in detail in~\cite{classification}. Here we want
only to concentrate on various notions of amenability.
\smallskip
Recall that $G$ is called \emph{coamenable} if $\|\Gamma_U\|=\dim H_U$
for every finite dimensional unitary representation $U$. There are a
number of equivalent conditions, but using this definition as our
starting point we immediately get that
$$
\Rep G \ \ \text{is amenable}\Leftrightarrow G\ \ \text{is coamenable,
and of Kac type}.
$$
Coamenability of $G$ is known to be equivalent to amenability of the
dual discrete quantum group~$\hat G$. Recall that the algebra of
bounded functions on $\hat G$ is defined by $\ell^\infty(\hat
G)=\ell^\infty\text{-}\oplus_{s\in\Irr(G)}B(H_s)$, and the coproduct
${\hat\Delta}\colon\ell^\infty(\hat G)\to\ell^\infty(\hat
G)\mathbin{\bar{\otimes}}\ell^\infty(\hat G)$ is defined by duality from the product
on $\mathbb{C}[G]$, if we view $\ell^\infty(\hat G)$ as a subspace of
$\mathbb{C}[G]^*$ by associating to a functional $\omega\in\mathbb{C}[G]^*$ the
collection of operators $\pi_s(\omega)=(\iota\otimes\omega)(U_s)\in
B(H_s)$, $s\in\Irr(G)$. The quantum group $\hat G$ is called amenable,
if there exists a right invariant mean on $\hat G$, that is, a state
$m$ on $\ell^\infty(\hat G)$ such that
$$
m(\iota\otimes\phi){\hat\Delta}=\phi(\cdot)1\ \ \text{for any normal linear
functional}\ \phi\ \text{on}\ \ell^\infty(\hat G).
$$
The restriction of such an invariant mean to $Z(\ell^\infty(\hat
G))\cong\ell^\infty(\Irr(G))$ defines a right invariant mean on the
fusion algebra of $\Rep G$ equipped with the quantum dimension
function. Therefore
$$
\hat G \ \ \text{is amenable}\Rightarrow \Rep G\ \ \text{is weakly
amenable}.
$$
Among various known characterizations of coamenability the implication
($\hat G$ is amenable $\Rightarrow$ $G$ is coamenable) is probably the
most nontrivial. This was proved independently
in~\cite{MR2276175}*{Theorem~3.8}
and~\cite{MR2113848}*{Corollary~9.6}. We will show now that our
results on amenable functors are generalizations of this.
\begin{theorem} If $\hat G$ is amenable, then the forgetful functor
$F\colon\Rep G\to\mathrm{Hilb}_f$ is amenable, and therefore $G$ is
coamenable.
\end{theorem}
\begin{proof} We will only consider the case when $\Irr(G)$ is at most
countable, so that $\Rep G$ satisfies our standing assumptions, the
general case can be easily deduced from this.
As discussed in \cite{arXiv:1310.4407}*{Section~4.1}, the space
$\hat{\mathcal{C}}(U, V)$ can be identified with the space of elements
$$
\eta\in \ell^\infty(\hat G)\otimes B(H_U,H_V)\ \ \text{such that}\ \
V^*_{31}(\alpha\otimes\iota)(\eta)U_{31}=1\otimes \eta,
$$
where $\alpha\colon \ell^\infty(\hat G)\to L^\infty(G)\bar\otimes
\ell^\infty(\hat G)$ is the left adjoint action of $G$. Under this
identification we have
$$
\iota_Y\otimes\eta=(\iota\otimes\pi_Y\otimes\iota)({\hat\Delta}\otimes\iota)(\eta),
$$
where $\pi_Y\colon\ell^\infty(\hat G)\to B(H_Y)$ is the representation
defined by $Y$, while the element $\eta\otimes\iota_Y$ has the obvious
meaning. From this we immediately see that if $m$ is a right invariant
mean on $\hat G$, then the maps $m\otimes\iota\colon \ell^\infty(\hat
G)\otimes B(H_U,H_V)\to B(H_U,H_V)$ define a right invariant mean for
$F$. Thus $F$ is amenable. By Theorem~\ref{tamnefunctor} we conclude
that $\|\Gamma_U\|=\dim F(U)=\dim H_U$ for every $U$, so $G$ is
coamenable. \end{proof}
\subsection{Subfactor theory}
Let $N\subset M$ be a finite index inclusion of II$_1$-factors. Denote
by $\tau$ the tracial state on $M$, and by $E$ the trace-preserving
conditional expectation $M\to N$. We denote $[M : N]=\Ind E$, and
the minimal index of $N \subset M$ by $[M : N]_0$. Put $M_{-1}=N$,
$M_0=M$, and choose a tunnel
$$
\dots\subset M_{-3}\subset M_{-2}\subset M_{-1}\subset M_0,
$$
so that $M_{-n+1}$ is the basic extension of $M_{-n-1}\subset M_{-n}$
for all $n\ge1$. For every $j\le 1$ denote by $M_j^\mathrm{st}\subset M_j$ the
$s^*$-closure of $\cup_{n\ge1}(M_{j-n}'\cap M_j)$ with respect to the
restriction of $\tau$. The inclusion $N^\mathrm{st}\subset M^\mathrm{st}$ of finite
von Neumann algebras is called a standard model of $N\subset
M$~\cite{MR1278111}.
Let $\mathcal{B}_N(M)$ be the full C$^*$-tensor subcategory of the category
$\mathrm{Hilb}_N$ of Hilbert bimodules over~$N$ generated by~$L^2(M)$.
Let $M_1$ be the basic extension of $N\subset M$, so that
$\Endd_{N\mhyph N}(L^2(M))\cong N'\cap M_1$. The embedding $N \to M_1$
induces a morphism $L^2(N) \to L^2(M) \otimes_N L^2(M)$ in $\mathcal{B}_N(M)$,
which defines a solution of the conjugate equations for $L^2(M)$ up to a scalar normalization. Moreover, it can be shown (compare with Proposition~\ref{pminindex}) that the categorical trace corresponds to the minimal conditional expectation $M_1\to N$, and consequently $d(L^2(M))=[M_1 : N]^{1/2}_0=[M : N]_0$. It is also known, see Proposition~\ref{pocneanu}, that the inductive system of the algebras
$\Endd_{N\mhyph N}(L^2(M)^{\otimes_Nn})$, with respect to the embeddings
$T\mapsto \iota_{L^2(M)}\otimes T$, can be identified with
$(M_{-2n+1}'\cap M_1)_{n\ge1}$ in such a way that the shift
endomorphism $T\mapsto T\otimes\iota_{L^2(M)}$ of
$\cup_{n\ge1}\Endd_{N\mhyph N}(L^2(M)^{\otimes_Nn})$ corresponds to the
endomorphism $\gamma^{-1}$ of $\cup_{n\ge1}(M_{-2n+1}'\cap M_1)$,
where $\gamma$ is the canonical shift.
The normalized categorical trace on $\Endd_{N\mhyph N}(L^2(M))$ defines a
probability measure~$\mu_\mathrm{st}$ on the set of isomorphism classes of
simple submodules of $L^2(M)$. More explicitly, it can be shown that
the value of the normalized categorical trace on any minimal
projection $p\in N'\cap M_1$ equals
$$
(\tau(p)\tau'(p))^{1/2}\frac{[M : N]}{[M : N]_0},
$$
where $\tau'$ is the unique tracial state on $N'\subset B(L^2(M))$. See~\cite{MR976765}*{Section~2}
and~\cite{MR1278111}*{Section~1.3.6} for related results. Then the
measure~$\mu_\mathrm{st}$ is defined~by
$$
\mu_\mathrm{st}([pL^2(M)])=m_p(\tau(p)\tau'(p))^{1/2}\frac{[M :
N]}{[M : N]_0},
$$
where $m_p$ is the multiplicity of $pL^2(M)$ in $L^2(M)$.
Recall that an inclusion for which $[M : N]=[M : N]_0$, is called
extremal. From the above considerations, unless $N\subset M$ is extremal,
we see that the categorical trace defines a tracial state of
$\cup_{n\ge1}(M_{-2n+1}'\cap M_1)$ that is different from~$\tau$.
\smallskip
Let us first review what our results say about $(\mathcal{B}_N(M),\mu_\mathrm{st})$ for extremal inclusions. From the identification of $\cup_{n\ge1}\Endd_{N\mhyph N}(L^2(M)^{\otimes_Nn})$ with $\cup_{n\ge1}(M_{-2n+1}'\cap M_1)$ we conclude that the von Neumann algebra $\mathcal{N}_{L^2(M)}$ constructed in Section~\ref{sec:longo-roberts-appr} is
isomorphic to $M_1^\mathrm{st}$. More precisely, we take $V=L^2(M)$ for the
construction of~$\mathcal{N}_{L^2(M)}$, so unless $N'\cap M_1$ is abelian, we
apply the modification of our construction of the algebras~$\mathcal{N}_U$
discussed in Remark~\ref{rmultiplicity}. The subalgebra
$\mathcal{N}\subset\mathcal{N}_{L^2(M)}$ corresponds then to $N^\mathrm{st}=\gamma^{-1}(M_1^\mathrm{st})\subset
M_1^\mathrm{st}$. In particular, $N^\mathrm{st}$ is a factor if and only if $\mu_\mathrm{st}$ is ergodic. Proposition~\ref{prop:p-bdry-as-rel-comm-LR0}
translates into the following statement, which is closely related to a
result of Izumi~\cite{MR2059809}.
\begin{proposition} \label{pIzumi} Let $\Pi\colon\mathcal{B}_N(M)\to\mathcal{P}$ be the Poisson boundary of $(\mathcal{B}_N(M),\mu_\mathrm{st})$. Then, assuming that $N\subset M$ is extremal, we have
$$
\mathcal{P}(L^2(M))\cong ({N^\mathrm{st}})'\cap M_1^\mathrm{st}.
$$
\end{proposition}
More generally, by the same argument we have
$\mathcal{P}(L^2(M)^{\otimes_Nn})\cong ({M_{-2n+1}^\mathrm{st}})'\cap M_1^\mathrm{st}$. Since
$L^2(M)$ contains a copy of the unit object induced by the inclusion
$N \to M$, we have $\mu_\mathrm{st}(e)>0$. Hence the supports of $\mu^n_\mathrm{st}$
are increasing, and therefore the isomorphisms
$\mathcal{P}(L^2(M)^{\otimes_Nn})\cong ({M_{-2n+1}^\mathrm{st}})'\cap M_1^\mathrm{st}$
completely describe the morphisms in the category $\mathcal{P}$. In fact,
recalling that $M_1^\mathrm{st}$ is the basic extension of $N^\mathrm{st}\subset
M^\mathrm{st}$, see ~\cite{MR1278111}*{Section~1.4.3}, we may conclude that
$\mathcal{P}$ can be identified with $\mathcal{B}_{N^\mathrm{st}}(M^\mathrm{st})$. We leave it to the
interested reader to find a good description of the functor $\Pi\colon
\mathcal{B}_N(M)\to\mathcal{B}_{N^\mathrm{st}}(M^\mathrm{st})$.
\smallskip
Consider the principal graph $\Gamma_{N,M}$ of $N\subset M$. Then
$\Gamma_{L^2(M)}$ can be identified with $\Gamma_{M,M_1}
\Gamma_{M,M_1}^t$. Recall also that we have the equality
$\|\Gamma_{N,M}\| = \|\Gamma_{M,M_1}\|$
by~\cite{MR1278111}*{Section~1.3.5}.
Turning now to Theorem~\ref{tminimal} and Proposition~\ref{pminindex},
we get the following result (again, to be more precise we use the
modification of the construction of~$\mathcal{N}_U$ described in
Remark~\ref{rmultiplicity}).
\begin{theorem} \label{tPopaext} Assume $N \subset M$ is extremal and
$N^\mathrm{st}$ is a factor. Then we have
$$
\|\Gamma_{N,M}\|^4=[M_1^\mathrm{st} : N^\mathrm{st}]_0.
$$
\end{theorem}
If $M^\mathrm{st}$ is also a factor, this can of course be formulated as
$\|\Gamma_{N,M}\|^2=[M^\mathrm{st} : N^\mathrm{st}]_0$.
\smallskip
Applying Theorem~\ref{tFKVR} we get the following result, which recovers part of Popa's characterization of extremal subfactors with strongly amenable standard invariant~\cite{MR1278111}*{Theorem~5.3.1}.
\begin{theorem}\label{thm:amen-equiv-high-rel-comm-std-mdl}
Assume $N \subset M$ is extremal. The following conditions are
equivalent:
\begin{enumerate}
\item $N^\mathrm{st}$ is a factor and $\|\Gamma_{N,M}\|^2=[M : N]_0$;
\item $(M^\mathrm{st}_{-2n+1})'\cap M_1^\mathrm{st}=M_{-2n+1}'\cap M_1$ for all
$n\ge1$.
\end{enumerate}
\end{theorem}
\begin{proof} As we already observed, the condition that $N^\mathrm{st}$ is a factor in (i) means exactly that the measure $\mu_\mathrm{st}$ is ergodic. The condition $\|\Gamma_{N,M}\|^2=[M : N]_0$ means that $\|\Gamma_{L^2(M)}\|=d(L^2(M))$. Since the module $L^2(M)$ is self-dual and generates $\mathcal{B}_N(M)$, this condition is equivalent to the amenability of $\mathcal{B}_N(M)$.
On the other hand, by Proposition~\ref{pIzumi} and its extension to
the modules $L^2(M)^{\otimes_Nn}$ discussed above, condition (ii) is
equivalent to triviality of the Poisson boundary of
$(\mathcal{B}_N(M),\mu_\mathrm{st})$.
This shows that the equivalence of (i) and (ii) is indeed a
consequence of Theorem~\ref{tFKVR}. \end{proof}
If we write the proof of the implication (ii)$\Rightarrow$(i) in terms
of the algebras $M_{-2n+1}'\cap M_1$ instead of
$\Endd_{N\mhyph N}(L^2(M)^{\otimes_Nn})$, we get an argument similar to Popa's
proof based on~\cite{MR1111570}, which was our inspiration. On the
other hand, our proof of (i)$\Rightarrow$(ii) seems to be very
different from Popa's arguments.
\smallskip
Next, let us comment on the nonextremal case. One possibility is to consider the completion of $\cup_{n\ge1}(M_{j-n}'\cap M_j)$ with respect to the trace induced by the minimal conditional expectation (that is, the categorical trace) instead of $\tau$. Then all the above statements continue to hold if we replace $N^\mathrm{st}$ and $M^\mathrm{st}$ by the corresponding new von Neumann
algebras. Note that the inclusion $N^\mathrm{st}\subset M^\mathrm{st}$ defined this way is the standard model in the
conventions of~\cite{MR1339767}. Then, for example, the implication (i)$\Rightarrow$(ii) in Theorem~\ref{thm:amen-equiv-high-rel-comm-std-mdl}
corresponds to~\cite{MR1339767}*{Lemma~5.2}.
But some results, notably~Theorem~\ref{tPopaext}, continue to hold for the
inclusion $N^\mathrm{st} \subset M^\mathrm{st}$ defined with respect to $\tau$ in the
nonextremal case also. The proof goes in basically the same way as in the extremal case, by noting that the proof of the inequality $\|\Gamma_U\|^2\ge [\mathcal{N}_U :\mathcal{N}]_0$ in Theorem~\ref{tminimal} did not depend on how exactly the inductive limit of the algebras $\mathcal{C}(V^{\otimes n}\otimes U)$ was completed to
get the factors $\mathcal{N}\subset\mathcal{N}_U$. Therefore we have
$$
\|\Gamma_{N,M}\|^4\ge [M_1^\mathrm{st} : N^\mathrm{st}]_0.
$$
The opposite inequality can be proved either by realizing that the
dimension function on $\mathcal{B}_{N^\mathrm{st}}(M_1^\mathrm{st})$ defines a dimension
function on $\mathcal{B}_N(M_1)$, or by the following string of
(in)equalities:
$$
\|\Gamma_{N,M}\|^4=\|\Gamma_{N,M_1}\|^2\le
\|\Gamma_{N^\mathrm{st},M_1^\mathrm{st}}\|^2 \le [M_1^\mathrm{st} : N^\mathrm{st}]_0,
$$
compare with~\cite{MR1278111}*{p.~235}. We remark that from this one
can easily obtain the implication (ii)$\Rightarrow$(vii)
in~\cite{MR1278111}*{Theorem~5.3.2} promised in~\cite{MR1278111}.
\bigskip
|
Subsets and Splits